Fact-checked by Grok 2 weeks ago

Diffraction

Diffraction is the bending of around the edges of an opening or an obstacle and the spreading out of beyond small openings, a fundamental characteristic of wave propagation observed in phenomena such as passing through slits, navigating corners, and water encountering barriers. This effect becomes prominent when the size of the obstacle or is comparable to the of the wave, leading to patterns that reveal the wave nature of the propagating . Diffraction occurs for all types of , including mechanical like and water , as well as electromagnetic such as and X-rays, and it underpins key experimental evidence for the wave-particle duality in . The historical development of diffraction theory traces back to early observations in the 17th century, with Italian physicist Francesco Grimaldi first documenting the bending of light around obstacles in 1665, though without a wave-based explanation at the time. In 1678, Dutch scientist Christiaan Huygens proposed his principle, stating that every point on a wavefront acts as a source of secondary spherical wavelets that propagate forward, providing a foundational framework for understanding wave propagation, reflection, and refraction, which later extended to diffraction. The modern wave theory of diffraction was advanced in 1818 by Augustin-Jean Fresnel, who applied Huygens's ideas to predict diffraction patterns mathematically, including the counterintuitive Poisson spot—a bright point at the center of a circular shadow—verified experimentally by Dominique Arago, thus supporting the wave model of light against prevailing particle theories. Further refinements came in the 19th century through the work of Gustav Kirchhoff in 1882, who derived diffraction from the wave equation using Green's theorem. At its core, diffraction is governed by the Huygens-Fresnel principle, which treats each point on an incoming wavefront as a secondary source of waves whose superposition determines the resulting intensity pattern, often expressed as I = I_0 \frac{\sin^2(N \phi / 2)}{\sin^2(\phi / 2)} for N slits separated by distance d, where \phi = 2\pi d \sin \theta / \lambda accounts for phase differences due to path length variations. Common manifestations include single-slit diffraction, producing a central bright band flanked by minima at angles \theta = m \lambda / a (where a is slit width and m is an integer), and multi-slit or grating diffraction, which enhances resolution through narrower principal maxima. These patterns arise from constructive and destructive interference, with the degree of spreading inversely proportional to the aperture size relative to the wavelength, limiting the precision of optical instruments like microscopes and telescopes via the diffraction limit \Delta \theta \approx 1.22 \lambda / D for circular apertures of diameter D. Diffraction plays a pivotal role in numerous scientific and technological applications, particularly in spectroscopy where diffraction gratings disperse light into its constituent wavelengths for atomic analysis, enabling the measurement of spectral lines with resolving power \lambda / \Delta \lambda = m N ( m being the order and N the number of slits). In X-ray crystallography, diffraction patterns from crystal lattices, first demonstrated by Max von Laue in 1912, allow determination of atomic structures essential for chemistry, biology, and materials science, as seen in the elucidation of DNA's double helix. Other uses include acoustic design, where diffraction influences sound propagation around barriers, and modern optics, such as in spectrometers that separate visible light spectra for astronomical observations.

History

Early Observations

The phenomenon of diffraction was first systematically observed and documented by the Jesuit and in the mid-17th century. In experiments conducted around 1660 and published posthumously in 1665 in his work Physico-mathesis de lumine, Grimaldi noted that passing through narrow slits or around the edges of opaque obstacles would spread out and produce colored bands beyond the geometric shadow, rather than strictly adhering to straight-line propagation. He coined the term "diffraction" (from the Latin diffractio, meaning "breaking apart") to describe this bending and dispersion, likening it to the way a stream of splits when encountering a thin obstacle. Building on such initial findings, early experiments in the late 17th and early 18th centuries further explored these and fringes. Scottish mathematician and astronomer James Gregory, in a 1673 letter to (secretary of the Royal Society), described observing spectral patterns produced by sunlight passing through the fine barbs of a bird feather, which acted as an early form of and revealed iridescent colors due to light's deviation around the structures. These observations highlighted the irregular bending of light near edges, though Gregory did not fully theorize the cause. Isaac Newton addressed these phenomena in his 1704 publication Opticks, where he referred to diffraction as "inflexions" of light rays. While Newton staunchly advocated a corpuscular (particle) theory of light and rejected wave explanations, he acknowledged the existence of colored fringes and patterns near obstacles or slits, describing experiments where light produced unexpected spectral displays beyond sharp shadows. In Book III and the appended Queries, Newton detailed how these inflexions generated rings and bands, attributing them to subtle deviations in ray paths without resolving their underlying mechanism, thus paving empirical groundwork for later wave-based interpretations. Diffraction-like effects were also recognized in natural atmospheric phenomena long before laboratory settings, with descriptions dating back to antiquity. Ancient observers, including in his (circa 350 BCE), documented luminous rings or halos encircling the , attributing them to interactions with atmospheric vapors, though modern understanding links these primarily to by suspended crystals in high-altitude clouds. These lunar halos, often appearing as 22-degree rings with subtle colored edges due to , arise when moonlight refracts through hexagonal ice prisms; related diffraction effects produce smaller around the through clouds containing tiny water droplets. Such events were routinely recorded in Chinese astronomical annals from the (481–221 BCE) onward, serving as omens or weather indicators.

Theoretical Development

In 1678, Dutch scientist proposed his principle, stating that every point on a wavefront acts as a source of secondary spherical wavelets that propagate forward, providing a foundational framework for understanding wave propagation, which later extended to diffraction. The theoretical development of diffraction began in the early with Thomas Young's double-slit experiment in 1801, which provided key evidence for the wave theory of light by demonstrating interference patterns, thereby linking wave propagation to diffraction effects. Young's work challenged the prevailing corpuscular model and laid the groundwork for understanding diffraction as a consequence of wave superposition. Building on this and Huygens' principle, advanced diffraction theory in his 1818 memoir submitted to the , where he modeled diffraction as the of secondary wavelets from wavefronts, earning a prize for his contributions. Fresnel's analysis predicted a bright at the center of a circular due to constructive , known as Poisson's spot, which initially seemed counterintuitive to proponents of the particle theory. In 1818, Dominique Arago experimentally confirmed this spot, providing empirical validation that solidified the wave model of diffraction. Concurrently in the 1810s, developed high-precision diffraction gratings using fine wires and ruled lines on glass, enabling the observation of lines in and establishing the far-field approximation now termed . His gratings, first constructed around 1821, allowed quantitative measurements of wavelengths and absorption features, transforming diffraction into a tool for . Further refinements came in the 19th century through the work of in 1882, who derived diffraction from the wave equation using . In the 20th century, diffraction theory extended to quantum realms with Louis de Broglie's 1924 hypothesis that particles exhibit wave-like properties, proposing matter waves with λ = h/p, which implied diffraction for electrons and other particles. This wave-particle duality was experimentally supported by diffraction patterns in electron beams. Further quantum formulation came through Richard Feynman's approach in the 1940s, which reinterpreted diffraction as the summation of probability amplitudes over all possible paths, unifying wave and particle descriptions in . A pivotal quantum application occurred in 1912 when demonstrated diffraction by crystals, revealing atomic lattice structures and confirming the wave nature of X-rays.

Fundamental Principles

Huygens-Fresnel Principle

The Huygens-Fresnel principle provides the foundational mechanism for understanding wave propagation and diffraction in . In 1678, proposed that every point on an advancing acts as a source of secondary spherical wavelets, which spread outward with the ; the new wavefront at any later time is the common tangent envelope to these wavelets. This construction reconciles the wave nature of light with observed phenomena like and , while contrasting sharply with geometric , where light travels in straight rays without bending. For instance, when a encounters an opaque obstacle, the secondary wavelets emanating from points along the unobstructed portions of the initial curve around the edges, allowing light to penetrate into shadowed regions through the superposition of these expanding spheres—thus explaining the bending of light that geometric cannot account for. In 1818, refined Huygens's idea by incorporating the principle of among the secondary wavelets, recognizing that their must combine constructively or destructively depending on path length differences to produce the observed intensity patterns. also introduced an obliquity factor, approximately (1 + [\cos](/page/Cos) \theta)/2, where \theta is the angle between the secondary wavelet's propagation direction and the normal to the initial , to model the observed decay in for wavelets propagating obliquely rather than perpendicularly. This factor ensures that contributions from secondary sources diminish appropriately with angle, enhancing the principle's predictive power for diffraction effects. Mathematically, the principle states that the disturbance ( and ) at an observation point P is determined by integrating the contributions from all secondary sources across the initial \Sigma: the total is proportional to \int_\Sigma \frac{1 + \cos \theta}{r} e^{i k r} \, [dS](/page/DS), where r is the distance from a source point on \Sigma to P, k = 2\pi / \lambda is the , and dS is the surface element—without deriving the full form here. This superposition propagates the wavefront forward, capturing how initial conditions evolve over distance. The principle assumes monochromatic waves of a single wavelength \lambda, as polychromatic light would require separate treatments for each frequency component to avoid phase complications. It also relies on far-field approximations for simplified calculations, though extensions handle near-field effects; these limitations ensure its applicability primarily to coherent, scalar wave scenarios in classical optics.

Interference and Superposition

The principle of superposition states that when multiple waves overlap in space, the total wavefield at any point is the vector sum of the individual wave amplitudes from each source. This linear addition applies to all wave phenomena, including , , and water waves, and is fundamental to understanding how complex wave patterns emerge from simpler components. In the context of interference, superposition leads to constructive interference when waves arrive in phase, meaning their crests and troughs align, resulting in maxima of intensity where amplitudes add up. Conversely, destructive interference occurs when waves are out of phase, such as when a crest meets a trough, leading to minima where amplitudes cancel out. These effects produce the alternating bright and dark fringes observed in interference patterns. Phase shifts arise from path length differences between waves, quantified by the phase difference δ = (2π/λ) Δx, where λ is the and Δx is the extra path length traveled by one wave relative to another. When Δx is an multiple of λ, the waves are in (δ = 2π n, n ), favoring constructive ; fractional multiples lead to partial or full cancellation. In diffraction, superposition of secondary wavelets—emanating from points across an or —interferes to form characteristic fringe patterns that extend beyond the geometric shadows expected from simple or . This redistributes into regions of reinforcement and cancellation, creating the bending and spreading of around edges. Diffraction inherently requires the nature of the phenomenon, as particle models predict straight-line trajectories without such spreading or effects, whereas propagate continuously and overlap to produce these observable deviations.

Mathematical Framework

Fraunhofer Diffraction

Fraunhofer diffraction describes the diffraction pattern observed in the far-field limit, where both the source and the observation point are effectively at infinite distance from the diffracting , resulting in plane wave propagation. This approximation simplifies the analysis by assuming that incoming and outgoing wavefronts are planar, which is valid when the observation distance z satisfies z \gg a^2 / \lambda, with a being the characteristic size of the and \lambda the of the . The formulation begins with the Huygens-Fresnel principle, which posits that the field at an observation point P is the superposition of secondary wavelets emanating from each point in the . The general diffraction from this principle is approximated in the far field by neglecting higher-order terms in the phase expansion. Specifically, the distance r from an aperture point (\xi, \eta) to P(x, y, z) is expanded as r \approx z + (x\xi + y\eta)/z, under the linear phase approximation, where the quadratic term ( \xi^2 + \eta^2 ) / (2z) is omitted. Angular coordinates are scaled such that \theta_x \approx x/z and \theta_y \approx y/z, leading to a that depends linearly on these angles. This yields the Fraunhofer diffraction for the field U(x, y) at the observation plane: U(x, y) = \frac{i}{\lambda z} e^{ikz} \iint A(\xi, \eta) \exp\left[ -i \frac{2\pi}{\lambda z} (x \xi + y \eta) \right] d\xi \, d\eta, where A(\xi, \eta) is the aperture function (complex amplitude transmittance), assuming monochromatic illumination and paraxial propagation. A key insight is that this integral represents the Fourier transform of the aperture function A(\xi, \eta), with spatial frequencies f_x = x / (\lambda z) and f_y = y / (\lambda z). Thus, the far-field diffraction pattern is the of the aperture transmittance, scaled by the observation geometry; the intensity pattern I(x, y) = |U(x, y)|^2 then follows from the squared modulus. This interpretation, central to , assumes monochromatic light and neglects obliquity factors for simplicity in the paraxial regime. This framework is foundational for analyzing patterns from simple apertures, such as the sinc distribution in single-slit diffraction or periodic intensity in gratings.

Fresnel Diffraction

Fresnel diffraction describes the bending of around obstacles or through when the observation point is in the near-field or , where the distance z from the aperture to is comparable to the square of the aperture dimension a divided by the \lambda, specifically z \lesssim a^2 / \lambda. This regime accounts for the of the wavefronts, unlike the far-field where spherical waves can be treated as plane waves. In this near-field setup, the diffraction pattern exhibits complex intensity variations due to the finite distance, including effects like the Poisson spot in circular . The amplitude of the diffracted U(P) at an observation point P(x, y, z) is calculated using the full Huygens-Fresnel diffraction , which integrates contributions from secondary wavelets across the : U(P) = \frac{1}{i\lambda} \iint \frac{A(\xi, \eta)}{r} \exp(ikr) \cos \chi \, d\xi \, d\eta, where A(\xi, \eta) is the , r = \sqrt{z^2 + (x - \xi)^2 + (y - \eta)^2} is the distance from a point (\xi, \eta, 0) on the to P, k = 2\pi / \lambda is the wave number, and \cos \chi is the obliquity factor approximating the directional dependence of the secondary sources, often taken as the cosine of the angle between the normal to the and the line to P. This captures the exact phase and amplitude variations without the quadratic phase approximation of the far . A key qualitative tool for understanding is the division of the into Fresnel zones, which are concentric half-period zones centered on the observation point's projection. Each zone corresponds to a region where the path length to P increases by \lambda/2 relative to the previous zone, resulting in alternating positive and negative contributions to the total due to the 180-degree phase shift between adjacent zones. For an unobstructed , the odd zones contribute constructively while even zones destructively interfere, leading to an intensity at P approximately one-quarter that of the first zone alone. This zonal construction predicts phenomena like the bright spot behind a circular obstacle. Fresnel zones also enable the design of focusing devices such as , constructed by alternately blocking or phase-shifting transparent and opaque rings corresponding to the zones, allowing only the positive-contributing zones to transmit and constructively interfere at a . For example, a transmission with alternating opaque and transparent annuli can focus a to a spot with a determined by the zone radii, offering a flat alternative to curved lenses for applications in and ; such zone plates achieve ~10% efficiency in the . Phase-optimized designs can reach up to 40% efficiency by modulating phase to reduce losses from destructive . Without deriving the full zone radii, such plates achieve focusing efficiencies up to 40% for optimized designs by reinforcing the first zone's . As the propagation distance z becomes very large compared to a^2 / \lambda, the wavefront curvature terms diminish, and the simplifies to the Fraunhofer form, where the diffraction pattern is the of the evaluated in the far field, neglecting the $1/r variation and higher-order phase terms. This transition highlights how encompasses the more general near-field behavior, while Fraunhofer provides a computationally simpler far-field .

Optical Diffraction Phenomena

Single-Slit Diffraction

Single-slit diffraction refers to the bending and spreading of light waves passing through a narrow rectangular , resulting in an pattern observable in the far field. This phenomenon is a fundamental demonstration of wave , where the slit acts as a secondary source of cylindrical wavefronts according to the Huygens-Fresnel principle, leading to constructive and destructive at different angles. The setup typically involves illuminating a slit of width a with a monochromatic of \lambda, such as from a coherent source, under Fraunhofer conditions where the observation distance is much larger than both the slit width and the wavelength divided by the angular spread. The intensity pattern arises from the coherent summation of wavelets emanating from infinitesimal elements across the slit. In the Fraunhofer approximation, valid for distant observation points or when using a focusing lens, the electric field at an angle \theta from the normal is given by the diffraction integral over the aperture: E(\theta) \propto \int_{-a/2}^{a/2} \exp\left(i \frac{2\pi y \sin\theta}{\lambda}\right) \, dy, where y is the coordinate along the slit width. This integral evaluates to a sinc function: E(\theta) \propto a \, \mathrm{sinc}\left( \frac{\pi a \sin\theta}{\lambda} \right), with the intensity I(\theta) = |E(\theta)|^2 proportional to I_0 \left[ \frac{\sin\beta}{\beta} \right]^2, where I_0 is the at \theta = 0 and \beta = \frac{\pi a \sin\theta}{\lambda} represents half the phase difference across the slit. The derivation assumes uniform illumination and neglects near-field effects, focusing on the far-field angular distribution. Dark minima in the pattern occur where the path differences cause complete destructive interference, specifically at angles satisfying a \sin\theta = m\lambda for integer m = \pm 1, \pm 2, \dots, corresponding to \beta = m\pi. The central maximum, centered at \theta = 0, has an angular full width of approximately $2\lambda / a, with subsequent maxima decreasing in intensity and becoming more pronounced for smaller slit widths relative to the wavelength. This scaling highlights how narrower slits produce broader diffraction patterns, emphasizing the wave nature of light. In experimental observations, a helium-neon laser (\lambda = 632.8 \, \mathrm{nm}) is directed through a precision slit mounted on an optical bench, with the diffraction pattern projected onto a screen several meters away or captured using a focusing lens at its focal plane for precise measurement. Photodetectors or CCD cameras scan the pattern to verify the predicted intensity profile, confirming the theoretical minima and the sinc-squared envelope. Such setups, often using slits with widths on the order of 0.02 to 0.1 mm, demonstrate the pattern's sensitivity to wavelength and aperture size.

Diffraction Grating

A is an optical component consisting of a periodic of numerous closely spaced slits or grooves that disperses light into its constituent s through and . This structure enables the production of well-defined spectra, making it essential for spectroscopic applications. Unlike a single slit, which produces a broad diffraction pattern, the grating's periodicity leads to reinforced principal maxima at specific angles, allowing for high-resolution wavelength separation. The standard setup for a transmission diffraction grating involves N parallel slits, each of width a, with centers spaced a d apart (where a < d), resulting in a total grating width of approximately N d. In the Fraunhofer approximation, valid for far-field conditions where the observation plane is at a large from the grating, the diffracted intensity pattern is derived by summing the contributions from each slit using the Huygens-Fresnel principle. The amplitude from the n-th slit includes a phase factor exp(i n ϕ), where ϕ = (2π/λ) d sinθ, leading to an interference factor of [sin(Nϕ/2)/sin(ϕ/2)]². This multi-slit term produces sharp principal maxima at angles θ satisfying sinθ = mλ/d, where m is an integer representing the diffraction order, modulated by the single-slit diffraction envelope, which has the form of a sinc² function centered at θ=0. For large N, the interference factor approximates a series of delta function peaks at the principal maxima locations, weighted by the envelope. The resolving power of a diffraction grating, defined as R = λ/Δλ (where Δλ is the smallest resolvable wavelength difference), is given by R = N m for the m-th order maximum, highlighting the grating's ability to distinguish closely spaced spectral lines through the narrowness of the principal peaks. Blazed gratings enhance efficiency by shaping the groove facets at a blaze angle θ_B to direct more light into a desired order, satisfying mλ_B = 2d sinθ_B for the blaze wavelength λ_B in reflection mode, achieving efficiencies up to 50% or more in the targeted order compared to uniform gratings. Diffraction gratings operate in either transmission mode, where light passes through a transparent substrate with etched grooves, or reflection mode, where light bounces off a ruled surface (often coated with a reflective material like aluminum), with the latter preferred for higher efficiency and compactness in many instruments.

Circular Aperture Diffraction

Circular aperture diffraction arises in optical systems with round openings, such as lenses and telescopes, where the radial symmetry of the aperture produces a characteristic diffraction pattern known as the . This pattern consists of a central bright spot surrounded by concentric rings of alternating intensity, resulting from the interference of light waves emanating from different points across the aperture. The phenomenon was first theoretically described by in 1835, who calculated the diffraction effects for a circular object-glass in the far-field approximation. The intensity distribution in the diffraction pattern for a circular aperture of radius a illuminated by monochromatic light of wavelength \lambda is given by I(\theta) = I_0 \left[ \frac{2 J_1(ka \sin \theta)}{ka \sin \theta} \right]^2, where I_0 is the intensity at the center (\theta = 0), J_1 is the first-order Bessel function of the first kind, k = 2\pi / \lambda is the wave number, and \theta is the angular displacement from the optical axis. This expression, derived from the Fraunhofer diffraction integral in polar coordinates, transforms the circular aperture function into a Bessel transform, yielding the radial symmetry of the pattern. The first minimum of this intensity occurs at \sin \theta \approx 1.22 \lambda / (2a), marking the edge of the central Airy disk and setting a fundamental limit on angular resolution in circularly symmetric optical systems. In contrast to the single-slit diffraction pattern, which features a rectangular aperture and produces a sinc-function intensity distribution with linear side lobes, the circular aperture's pattern exhibits rotational symmetry and circular rings due to the azimuthal integration in the diffraction integral. This difference highlights how aperture geometry influences the resulting interference structure, with the Airy pattern's compact central disk providing superior resolution for point sources in imaging applications compared to the broader sinc envelope.

General Aperture Effects

In the Fraunhofer diffraction regime, the far-field diffraction pattern produced by an arbitrary aperture is given by the two-dimensional Fourier transform of the aperture's complex amplitude transmittance function, which describes how the incident wavefront is modulated by the aperture. This approach allows for the analysis of non-standard geometries by representing the aperture as a function t(x, y), where the diffracted field in the observation plane is proportional to U(f_x, f_y) = \iint_{-\infty}^{\infty} t(x, y) \exp\left[-i 2\pi (f_x x + f_y y)\right] \, dx \, dy, with f_x and f_y denoting spatial frequencies corresponding to angular coordinates. For straight-edged apertures, this integral can often be evaluated in closed form using geometric properties. Babinet's principle provides a useful relation for complementary apertures, stating that the diffracted fields from an opaque screen and its complementary aperture sum to the unobstructed incident wave field, enabling efficient computation of one pattern from the other. This principle holds under scalar diffraction approximations and is particularly valuable for irregular shapes where direct calculation is complex. Computational simulation of these diffraction patterns frequently employs the fast Fourier transform (FFT) algorithm to efficiently evaluate the integral for discrete aperture representations, making it practical to model propagation for arbitrary geometries on digital computers. The aperture shape significantly influences the resulting pattern's symmetry; for instance, a rectangular aperture yields a separable, symmetric sinc-like pattern, while a triangular aperture introduces asymmetry, producing directional lobes and deformed intensity distributions due to the non-uniform boundary contributions in the Fourier domain. More generally, the pupil function P(\xi, \eta) encapsulates both amplitude and phase variations across the aperture, such as those induced by apodization or aberrations, with the diffraction pattern emerging as its Fourier transform; this formulation extends the basic model to account for realistic optical elements beyond ideal binary transmittance. For specific symmetric cases like the circular aperture, the pattern exhibits radial symmetry, but the general framework reveals how deviations in shape or pupil properties lead to anisotropic spreading.

Advanced Diffraction Effects

Laser Beam Propagation

In laser beam propagation, diffraction fundamentally limits the ability to maintain a collimated beam over long distances, causing inevitable spreading due to the wave nature of light. For coherent laser sources, the represents the fundamental transverse mode that minimizes this diffraction-induced divergence while satisfying the in free space. This mode arises as an exact solution to the under the paraxial approximation, enabling self-similar propagation where the beam profile scales predictably with distance. The electric field of a Gaussian beam propagating along the z-axis can be expressed as
E(r, z) = E_0 \frac{w_0}{w(z)} \exp\left[-\frac{r^2}{w(z)^2}\right] \exp\left[i(kz + \phi)\right],
where E_0 is the amplitude at the beam waist, w_0 is the waist radius (defined at $1/e^2 intensity), w(z) is the beam radius at axial distance z, r is the radial coordinate, k = 2\pi/\lambda is the wavenumber, and \phi accounts for phase terms including the Gouy phase and curvature. The beam radius evolves as w(z) = w_0 \sqrt{1 + (z/z_R)^2}, where z_R is the Rayleigh range, marking the transition from the near field (collimated region) to the far field (diverging region).
Diffraction spreading in the far field is characterized by the half-angle divergence \theta = \lambda / (\pi w_0), which quantifies the asymptotic conical expansion of the beam; for example, a helium-neon laser beam with w_0 = 0.5 mm at \lambda = 633 nm exhibits \theta \approx 0.4 mrad. The Rayleigh range is given by z_R = \pi w_0^2 / \lambda, defining the distance over which the beam area doubles; within z_R, diffraction effects are minimal, while beyond it, the beam behaves as a spherical wave originating from the waist. These parameters ensure that Gaussian beams achieve the lowest possible beam parameter product M^2 = 1, representing diffraction-limited performance. The self-similar nature of Gaussian beam propagation is elegantly described using the ABCD ray transfer matrix formalism under the paraxial approximation, which tracks transformations of the complex beam parameter q(z) = z + i z_R through optical systems. For free-space propagation over distance d, the ABCD matrix is \begin{pmatrix} 1 & d \\ 0 & 1 \end{pmatrix}, yielding q(z + d) = (A q(z) + B) / (C q(z) + D), preserving the Gaussian form without distortion. This method, originally developed for lenslike media, extends to vacuum propagation and underpins the design of laser resonators and beam delivery systems. In contrast to a uniform aperture illumination, which produces diffraction patterns with sidelobes (e.g., sinc for slits or Airy disks for circles) and angular spreading \theta \approx \lambda / D where D is the aperture diameter, Gaussian beams exhibit smoother, sidelobe-free profiles with equivalent far-field divergence but superior uniformity. The absence of sharp edges in the Gaussian intensity distribution (\propto \exp[-2r^2/w^2]) reduces unwanted diffraction artifacts, making it the preferred mode for applications requiring stable, long-distance propagation.

Diffraction-Limited Imaging

In optical imaging systems, diffraction imposes a fundamental limit on the achievable resolution, preventing the formation of a perfect point image from an ideal point source. This phenomenon arises because light waves passing through an aperture interfere constructively and destructively, spreading the image into a finite pattern rather than converging to a delta function. first articulated this diffraction limit in 1873 while developing the theory of microscopic image formation, establishing that the resolution in traditional optical microscopy cannot exceed approximately 0.2 micrometers for visible light, as finer details become blurred due to wave nature. The point spread function (PSF) quantifies this blurring in diffraction-limited systems, representing the three-dimensional intensity distribution produced by an infinitely small point source after passing through the optical system. In an ideal aberration-free lens, the PSF takes the form of an Airy pattern for a circular aperture, consisting of a bright central disk surrounded by concentric faint rings due to . This pattern describes how the system responds to a point object, with the overall image being the convolution of the true object with the PSF; thus, features smaller than the PSF width cannot be resolved distinctly. The minimum resolvable distance in such systems, known as the diffraction limit, is determined by the radius of the Airy disk, given by \delta = 0.61 \lambda / \mathrm{NA}, where \lambda is the wavelength of light and NA is the numerical aperture of the objective. This formula, derived from the , specifies the separation at which two point sources produce Airy disks just touching, with their combined intensity showing a detectable dip at the midpoint. For example, in a microscope using green light (\lambda \approx 550 nm) and an objective with NA = 1.4, \delta approaches 0.24 \mum, setting the scale for resolving cellular structures. In aberration-free systems, the Strehl ratio—a measure of peak intensity relative to the ideal diffraction-limited case—equals 1, indicating maximal concentration of light in the PSF core and optimal performance. To mitigate the broad sidelobes of the Airy PSF and sharpen the central peak for improved resolution of closely spaced points, apodization techniques modify the pupil function by applying amplitude or phase masks that taper the aperture transmission. For instance, a parabolic apodizer redistributes light intensity across the pupil, reducing the focal spot size and enhancing the dip between overlapping PSFs, though at the expense of decreased light throughput (diffraction efficiency often below 90%) and a narrower field of view. These methods trade overall brightness for contrast in high-resolution applications like microscopy, where preserving signal is critical.

Speckle Patterns

Speckle patterns arise from the coherent superposition of numerous scattered wavelets, each carrying a random phase due to diffuse reflection or scattering from a rough surface. When coherent light illuminates an optically rough surface, the reflected or transmitted wavefront is decomposed into many secondary wavelets with uncorrelated phases, leading to random interference that produces a granular intensity distribution known as speckle. This phenomenon is fundamentally tied to the coherence of the illuminating light, which ensures the phase relationships necessary for constructive and destructive interference. In fully developed speckle patterns, the intensity fluctuations follow a negative exponential probability distribution, given by p(I) = \frac{1}{\langle I \rangle} \exp\left( -\frac{I}{\langle I \rangle} \right), where \langle I \rangle is the mean intensity and also equals the variance \sigma^2. This distribution reflects the random nature of the phasors summing to form the field amplitude, resulting in Rayleigh-distributed amplitude and thus exponential intensity statistics. The contrast of a speckle pattern, defined as the standard deviation of intensity divided by the mean, equals 1 for fully developed, fully coherent cases, indicating maximum granularity. Partial coherence or superposition of multiple uncorrelated speckle patterns reduces this contrast, for example, to $1 / \sqrt{M} when adding M independent intensity patterns. Speckle patterns are classified into objective and subjective types based on their formation geometry. Objective speckle forms directly in the far-field or observation plane from scattering by a diffuse object under coherent illumination, independent of any imaging system. Subjective speckle, in contrast, appears in the image plane or on the retina when viewing the scattered light through a lens or optical system, resulting from the interference filtered by the aperture. In physics and metrology, the autocorrelation function of the speckle field provides insight into surface roughness, as its width correlates with the scatterer size or roughness scale on the diffuse surface. For instance, the decorrelation length of the intensity autocorrelation decreases with increasing surface roughness, enabling non-contact measurements of parameters like root-mean-square height. This property underpins applications such as roughness profiling, where shifts in speckle autocorrelation under varying illumination angles quantify surface texture without physical contact.

Babinet's Principle

Babinet's principle, formulated by French physicist in 1837, establishes a fundamental complementarity in wave patterns between an aperture in an opaque screen and its complementary obstacle. Babinet drew an analogy to the interference colors observed in thin soap films to illustrate how diffracted waves from complementary structures produce equivalent effects, emphasizing the principle's intuitive basis in wave superposition. The principle states that the diffraction pattern produced by an aperture is identical to that from its complementary opaque obstacle, except at the forward direction where the obstacle blocks the direct beam. Mathematically, this arises from the linearity of the wave equation: the total field U in the observation plane satisfies U_{\text{aperture}} + U_{\text{obstacle}} = U_{\text{plane}}, where U_{\text{plane}} is the unobstructed incident wave field, and the diffracted fields from the aperture and obstacle are complementary such that their sum reconstructs the free propagation. This relation holds because the boundary conditions for the two screens differ only in the sign of the field over the complementary region, leading to diffracted amplitudes that differ by the incident wave itself. A classic example is the diffraction from a narrow slit compared to that from a thin wire of the same width: the intensity patterns match everywhere except the central maximum, where the wire's pattern is the free wave minus the slit's. The principle applies equally to both (far-field) and (near-field) diffraction regimes, as the underlying linearity is independent of the propagation distance. Limitations include its neglect of polarization effects in electromagnetic waves, where vectorial nature requires modifications like Booker's duality for full validity, and assumptions of no multiple reflections or absorption at edges. It also fails for highly asymmetric complements or when surface interactions (e.g., in matter waves) introduce additional scattering.

Knife-Edge Diffraction

Knife-edge diffraction refers to the bending of waves around a sharp, semi-infinite plane obstacle, representing a canonical boundary case in wave propagation studies. This phenomenon occurs when an incident wave encounters an opaque half-plane, leading to interference patterns both in the geometrically illuminated region and the shadow. The effect is prominent in the near-field Fresnel regime, where the obstacle blocks part of the wavefront, allowing secondary wavelets from the edge to contribute to the field distribution. In the Fresnel approximation, the diffracted field amplitude U(v) at a point in the observation plane is proportional to the complex Fresnel integral: U(v) \propto \int_{-\infty}^{v} \exp\left(i \frac{\pi u^{2}}{2}\right) \, du, where v = x \sqrt{\frac{2}{\lambda z}} is the normalized coordinate, with x the transverse position, \lambda the wavelength, and z the propagation distance. This integral traces a path along the in the complex plane, where the real part is the Fresnel cosine integral C(v) = \int_{0}^{v} \cos\left(\frac{\pi t^{2}}{2}\right) \, dt and the imaginary part is the Fresnel sine integral S(v) = \int_{0}^{v} \sin\left(\frac{\pi t^{2}}{2}\right) \, dt, yielding U(v) = \frac{1 + i}{2} + \left[ C(v) + i S(v) \right]. The intensity is I(v) = \frac{|U(v)|^2}{2}, transitioning smoothly from oscillatory fringes in the illuminated region (v > 0) to decaying oscillations in the shadow (v < 0), with the boundary at v = 0 exhibiting I = 0.25 relative to the unobstructed field. A notable feature is the emergence of a bright fringe within the geometric shadow near the edge, arising from constructive interference of contributions from successive Fresnel zones partially exposed by the obstacle. These zones, concentric half-period regions of the wavefront, alternate in phase, but the edge allows an imbalance that produces intensity maxima just beyond the boundary, with the first peak at approximately v = -1.22 where I \approx 0.17. Intensity oscillations persist deeper into the shadow, damping exponentially as higher zones contribute less due to the $1/\sqrt{r} cylindrical wave decay from the edge. For high-frequency scenarios where the Fresnel number is large, the Geometric Theory of Diffraction (GTD), formulated by J. B. Keller in 1962, provides an asymptotic ray-based approximation. GTD extends geometrical optics by introducing diffracted rays that originate from the knife edge upon grazing incidence, propagating as cylindrical waves with amplitude scaling as (kr)^{-1/2} \exp(ikr), where k = 2\pi/\lambda and r is the distance from the edge. The diffraction coefficient for a half-plane incorporates angular dependence, such as D \propto -\frac{\exp(-i\pi/4)}{\sqrt{2\pi k} \sin(\phi/2)}, capturing shadow boundary discontinuities and multiple interactions in complex geometries. This approach simplifies computations for sharp features, yielding accurate field predictions outside transition regions. In radio wave propagation, knife-edge diffraction models signal loss over terrain horizons, such as hills or buildings, enabling beyond-line-of-sight communication. The International Telecommunication Union (ITU) recommends using the Fresnel parameter \nu = h \sqrt{\frac{2(d_1 + d_2)}{\lambda d_1 d_2}}, where h is the obstacle height and d_1, d_2 are distances to transmitter and receiver, to compute diffraction loss J(\nu) = 6.9 + 20 \log_{10}(\sqrt{\nu^2 + 1} + \nu) dB for \nu > -0.78. This results in oscillatory attenuation near the horizon, with losses typically 6-20 dB beyond the geometric line-of-sight, facilitating VHF/UHF coverage predictions in irregular terrains.

Diffraction Patterns and Analysis

Intensity Distributions

In diffraction patterns, the describes the spatial variation of across the observation plane, typically expressed in the far field as I(\theta, \phi) = |U(\theta, \phi)|^2, where U(\theta, \phi) represents the derived from the 's field via methods. This formulation arises from the scalar diffraction theory, where the squared captures constructive and destructive , producing a characteristic central bright lobe flanked by weaker that decay with angular distance from the center. The presence of is a universal feature of finite , resulting from the finite extent of the , and their relative intensities depend on the geometry, with higher indicating broader secondary maxima. The overall scale of the pattern exhibits an inverse proportionality to the aperture's linear dimension; for instance, the angular width of the central lobe \theta scales as \theta \approx \lambda / D, where \lambda is the and D is the size, such that larger apertures yield narrower, more concentrated patterns. This scaling principle, fundamental to in optical systems, ensures that diffraction effects become more pronounced for smaller apertures relative to the wavelength, compressing the pattern's extent in angle space. Consequently, in practical applications like or telescopes, optimizing aperture size balances against unwanted broadening of the . For systems involving high numerical apertures (NA > 0.5), polarization introduces vectorial effects that alter the scalar intensity predictions, necessitating vector diffraction theory to model the non-uniform electric field components across the wavefront. In such cases, the intensity distribution varies with the incident polarization state—linear, circular, or radial—leading to asymmetries in lobe shapes and enhanced sidelobe intensities due to depolarization at oblique angles. These effects are particularly critical in tightly focused beams, where the vectorial nature significantly differs from scalar approximations. Numerical evaluation of intensity distributions in simulations requires careful attention to sampling to avoid artifacts; adherence to the Nyquist criterion, demanding at least two samples per wavelength in the aperture plane, ensures faithful reproduction of lobes and sidelobes without aliasing. Insufficient sampling can artificially broaden sidelobes or introduce spurious peaks, compromising accuracy in computational optics. Qualitatively, patterns from rectangular apertures exhibit a broad central lobe tapering into symmetric sidelobes reminiscent of a sinc function, with secondary maxima decreasing gradually, whereas circular apertures produce the Airy pattern: a prominent central disk encircled by faint, concentric rings that diminish rapidly in intensity. These archetypal forms highlight how aperture shape influences the texture of the intensity landscape, with the Airy pattern's rings offering tighter confinement ideal for imaging applications.

Angular Dependence

The angular dependence of diffraction patterns arises primarily from the interference of wavelets emanating from different parts of an aperture or obstacle, with the spread governed by the ratio of wavelength to aperture dimension. In the far-field regime, known as Fraunhofer diffraction, the angular half-width to the first minimum for a single slit of width a is approximately \theta \approx \lambda / a under the small-angle approximation where \sin \theta \approx \theta. This approximation simplifies calculations when the diffraction angle is small, typically valid for \theta \ll 1 radian, and highlights how the pattern's angular extent scales inversely with aperture size. Wavelength plays a central role in determining the angular spread, as longer wavelengths produce greater diffraction angles for a fixed . For example, with \lambda \approx 700 nm diffracts more widely than with \lambda \approx 400 nm through the same , leading to broader patterns and reduced in systems. In diffraction gratings, this wavelength sensitivity manifests as angular dispersion, quantified by the relation \frac{d\theta}{d\lambda} = \frac{m}{d \cos \theta}, where m is the diffraction order and d is the grating groove spacing; higher orders or smaller d enhance the separation of wavelengths by angle. The nature of angular broadening differs markedly between near-field (Fresnel) and far-field regimes. In the near field, close to the (Fresnel number F \gg 1), the diffraction pattern evolves with propagation , exhibiting limited spread that retains details of the shape. Conversely, in the far field (F \ll 1, typically at distances z \gg a^2 / \lambda), the pattern stabilizes into a fixed independent of further , resembling the of the function with pronounced broadening. This transition underscores the importance of observation in perceiving effects. Diffraction angles become practically unobservable when the wavelength is much smaller than the or size, as the angular deviation \theta \propto \lambda / D (with D the dimension) approaches zero. For instance, visible light (\lambda \sim 500 nm) around everyday obstacles like buildings (D \sim 10 m) yields \theta \sim 5 \times 10^{-8} radians, indistinguishable from geometric . Observable diffraction requires comparable scales, where nature dominates over ray-like propagation.

Diffraction in Matter Waves

Electron Diffraction

Electron diffraction refers to the wave-like patterns produced when a beam of electrons interacts with a crystalline sample, demonstrating the de Broglie hypothesis that particles possess wave properties. The de Broglie wavelength \lambda of an electron is given by \lambda = h / p, where h is Planck's constant and p is the electron's . For non-relativistic electrons accelerated through a potential difference V (in volts), the p = \sqrt{2 m e V}, where m is the and e is the , yielding \lambda \approx 1.23 / \sqrt{V} nm. This short wavelength, typically on the order of 0.01–0.1 nm for accelerating voltages of 100–10,000 V, enables high-resolution probing of atomic-scale structures in solids. The experimental confirmation of wave nature came from the Davisson-Germer experiment in , where electrons accelerated to 54 eV were scattered off a surface, producing intensity peaks at specific angles consistent with Bragg diffraction of waves with \lambda \approx 0.165 nm. These peaks matched the expected diffraction from the (111) planes of face-centered cubic , providing direct evidence for de Broglie's matter waves and marking a foundational validation of wave-particle duality. In modern transmission electron diffraction (TED), a collimated electron beam passes through a thin sample in a transmission electron microscope, generating diffraction patterns on a detector. For polycrystalline samples, the random orientations of crystallites produce concentric rings in the diffraction pattern, reflecting averaged contributions from many grains. In contrast, single-crystal samples oriented along a zone axis yield discrete spot patterns, where each spot corresponds to a specific set of diffracting planes. These patterns arise from the three-dimensional Laue diffraction condition, expressed as \vec{a} \cdot (\vec{k} - \vec{k_0}) = 2\pi h, \vec{b} \cdot (\vec{k} - \vec{k_0}) = 2\pi k, and \vec{c} \cdot (\vec{k} - \vec{k_0}) = 2\pi l, where \vec{a}, \vec{b}, \vec{c} are the real lattice vectors, \vec{k_0} and \vec{k} are the incident and scattered wave vectors, and h, k, l are integers (Miller indices). TED is integral to transmission electron microscopy, where diffraction patterns complement high-resolution imaging by providing crystallographic information such as lattice parameters, phase identification, and orientation mapping, essential for materials characterization at the nanoscale.

Neutron Diffraction

Neutron diffraction utilizes thermal neutrons, which have de Broglie wavelengths typically ranging from 0.1 to 1 nm, suitable for probing atomic-scale structures in materials. The wavelength is given by the formula \lambda = \frac{h}{\sqrt{2 m E}}, where h is Planck's constant, m is the neutron mass, and E is the kinetic energy; for reactor sources, thermal neutrons at around 25 meV yield wavelengths near 0.18 nm, while cold neutron sources extend this to longer values for enhanced resolution in certain applications. These wavelengths arise from neutrons moderated to thermal energies in nuclear reactors, enabling diffraction patterns analogous to those from X-rays but with distinct interaction properties. The technique originated in the 1940s amid the Manhattan Project's nuclear research, with pioneering experiments at . Ernest O. Wollan conducted the first successful diffraction observations in December 1944 using the on crystals like and NaCl, demonstrating Bragg scattering. In 1946, Clifford G. Shull joined Wollan, and together they developed the first powder diffractometer in 1948, capturing polycrystalline diffraction patterns from NaCl and establishing the method's viability for structural analysis. Their work, conducted under constrained wartime conditions, laid the foundation for scattering as a structural tool, earning Shull, shared with Bertram N. Brockhouse, the 1994 . A key advantage of neutron diffraction is its deep penetration into matter due to the of neutrons with electrons, allowing non-destructive analysis of samples up to centimeters thick, unlike more surface-sensitive methods. Neutrons are particularly sensitive to isotopic differences, as scattering lengths vary significantly between isotopes (e.g., higher for than ), enabling isotopic substitution studies via nuclear form factors. Additionally, neutrons interact with magnetic moments through magnetic form factors, facilitating the determination of magnetic structures in materials like antiferromagnets. In powder samples, diffraction produces Debye-Scherrer rings, where scattered s form conical patterns intersected by detectors to yield intensity versus angle profiles for parameter refinement. These rings arise from the random orientations of crystallites, providing averaged structural information. In biological applications, diffraction excels at locating positions in proteins and enzymes, revealing bonding networks and states critical for function, as demonstrated in studies of rubredoxin and crystals. This capability stems from s' strong scattering from (via incoherent scattering) and deuterium labeling, offering insights unattainable with other diffraction techniques.

Crystal and Bragg Diffraction

Bragg's Law

Bragg's law describes the condition for constructive interference of waves diffracted by successive atomic planes in a periodic crystal lattice, arising from the path length difference between scattered waves. Consider a plane wave incident on a set of parallel crystal planes separated by distance d. The angle between the incident beam and the planes is \theta, and the scattered waves are observed at the same angle \theta (specular reflection). The extra path length traveled by the wave reflecting from the second plane compared to the first is $2d \sin \theta. For constructive interference, this path difference must equal an integer multiple of the wavelength \lambda, yielding the relation $2d \sin \theta = n \lambda, where n is a positive integer denoting the diffraction order. This scalar formulation assumes reflection from a one-dimensional stack of planes but extends to three-dimensional lattices through the concept of the , which maps the periodic structure into a space of vectors. The vectors \mathbf{G}_{hkl} are defined as \mathbf{G}_{hkl} = h \mathbf{b}_1 + k \mathbf{b}_2 + l \mathbf{b}_3, where h, k, l are integers () and \mathbf{b}_i are the basis vectors reciprocal to the real-space lattice vectors. Diffraction occurs only for wavevectors satisfying the Laue condition in vector form: the change in the wavevector \Delta \mathbf{k} = \mathbf{k}_\text{out} - \mathbf{k}_\text{in} = \mathbf{G}, where |\mathbf{k}_\text{in}| = |\mathbf{k}_\text{out}| = 2\pi / \lambda. This ensures the scattered wave is in phase across the lattice. The Ewald sphere construction visualizes allowed reflections in reciprocal space. Place the origin of the reciprocal lattice at the tail of the incident wavevector \mathbf{k}_\text{in} (of length $2\pi / \lambda) on a sphere of radius $2\pi / \lambda centered at the crystal position. As the crystal is rotated, reciprocal lattice points trace paths; a reflection is observed when a point \mathbf{G} intersects the sphere's surface, satisfying |\mathbf{k}_\text{in} + \mathbf{G}| = |\mathbf{k}_\text{in}| and thus the Laue condition. This geometric method highlights that not all \mathbf{G} contribute for a given wavelength and orientation, limiting observable reflections. The provide an equivalent scalar description for a general : \mathbf{a}_i \cdot (\mathbf{s} - \mathbf{s}_0) = h_i \lambda for i=1,2,3, where \mathbf{a}_i are the real-space vectors, \mathbf{s}_0 and \mathbf{s} are unit vectors along the incident and scattered directions, \lambda is the , and h_i are integers. These reduce to for planes normal to one axis. However, even when the Laue condition holds, the diffracted intensity is modulated by the F_{hkl}, which accounts for from atoms within the unit cell: F_{hkl} = \sum_j f_j \exp \left[ 2\pi i (h x_j + k y_j + l z_j) \right], where the sum is over all atoms j in the unit cell, f_j is the atomic scattering factor (approximately the number of electrons for X-rays), and (x_j, y_j, z_j) are . The intensity is proportional to |F_{hkl}|^2. Certain symmetries cause F_{hkl} = 0 for specific hkl, leading to extinction rules or systematic absences that reveal the . For example, in a centered , reflections vanish if h + k + l is odd, as phase differences cancel contributions from equivalent atoms. These rules arise directly from the exponential terms in the sum, providing constraints on possible atomic arrangements without solving the full phase problem.

X-Ray Crystallography

X-ray crystallography originated with Max von Laue's 1912 experiment, in which he demonstrated that X-rays passing through a crystal of zinc blende produced discrete diffraction spots, confirming the wave nature of X-rays and the periodic arrangement of atoms in crystals. In 1913, and his son William Lawrence Bragg applied their newly formulated to analyze these patterns, determining the structures of simple crystals like and by measuring reflection intensities. For their foundational contributions, Laue was awarded the 1914 , while the Braggs shared the 1915 prize, recognizing the birth of the field as a tool for atomic-scale structure determination. Key experimental techniques in X-ray crystallography include the rotating crystal method, introduced in the early 1910s, which involves mounting a on a rotating to expose multiple orientations to the X-ray beam, producing a comprehensive set of diffraction spots on or detectors. For polycrystalline materials, the powder diffraction method, developed by and in 1916, grinds samples into fine powders to achieve random orientations, yielding concentric Debye-Scherrer rings that represent averaged diffraction from many crystallites. Data analysis begins with processing diffraction intensities to compute maps via Fourier transforms, but the phase problem—requiring unknown phases of structure factors—necessitates specialized solutions. The Patterson function, introduced by Arthur Lindo Patterson in 1935, circumvents phases by transforming squared intensities into a map of interatomic vectors, facilitating initial heavy-atom positioning. Phases are then solved using isomorphous replacement, which introduces heavy atoms to create derivative crystals with measurable phase shifts, or direct methods, which probabilistically estimate phases from intensity statistics. The achievable , defined as the minimum interplanar spacing d_{\min} \approx \frac{[\lambda](/page/Lambda)}{2 \sin \theta_{\max}} where [\lambda](/page/Lambda) is the wavelength and \theta_{\max} the maximum scattering angle, typically reaches about 1.5 for atomic detail. By the 1950s, advances in these methods enabled the first determinations of protein structures, such as at 2 by et al. in 1960, revealing the polypeptide chain's three-dimensional fold and paving the way for understanding biomolecular function.

Coherence Requirements

Spatial Coherence

Spatial , also known as transverse , refers to the correlation of the of a wave across its in the direction perpendicular to propagation, which is essential for the formation of clear diffraction patterns over extended . In diffraction experiments, such as those involving slits or gratings, sufficient spatial ensures that the maintains a consistent relationship across the aperture, allowing to produce sharp fringes. Without adequate spatial , the diffraction pattern blurs due to phase variations, reducing contrast and . The Van Cittert-Zernike theorem provides a fundamental description of spatial from an incoherent extended source, stating that the mutual function between two points in the observation plane is the normalized of the source's intensity distribution. This theorem, originally derived by van Cittert and extended by Zernike, implies that for a source at a distance z from the observation plane, the degree of spatial decreases with separation from the , leading to a characteristic transverse coherence length l_c = \frac{\lambda z}{\pi \sigma}, where \lambda is the wavelength and \sigma is the root-mean-square size of a Gaussian source profile. For non-Gaussian sources, the coherence length scales inversely with the source's angular extent, highlighting how larger sources reduce .90042-7) In practice, extended sources like lamps produce low spatial , causing diffraction patterns to blur as phase differences across the average out, particularly affecting fine-scale features. To achieve high spatial , optical setups often employ pinholes or for spatial filtering, which select a smaller effective source size and enhance the , enabling observable diffraction from larger . The impact of insufficient spatial is evident in the washing out of high-frequency fringes in patterns, limiting the ability to resolve small angular structures. Spatial is commonly measured using Young's double-slit experiment, where the of fringes—defined as the contrast between maximum and minimum —directly quantifies the degree of between the slits. For slit separation d much smaller than the , fringes are sharp with visibility near unity; as d approaches or exceeds l_c, visibility drops, confirming the theorem's predictions. This method underscores the requirement for spatial in diffraction-based and .

Temporal Coherence

Temporal coherence in diffraction refers to the longitudinal of the field along the direction of , which is crucial for maintaining sharp patterns in polychromatic sources. It quantifies how well the relationship between different components persists over time, directly impacting the of diffraction fringes. For with a finite spectral , temporal incoherence leads to blurring or of the diffraction pattern, as contributions from wavelengths separated by Δλ interfere destructively beyond a certain path difference. The τ_c, defined as the duration over which the remains predictable, is inversely proportional to the Δν of the source: τ_c = 1/Δν. This time scale determines the maximum path difference for which can occur coherently. The corresponding l_c, the distance travels in that time, is l_c = c τ_c = c / Δν, where c is the ; equivalently, in terms of , l_c ≈ λ² / Δλ, with λ the central . These relations arise from the relationship between the temporal of and its power spectrum. In diffraction experiments, the quasi-monochromatic approximation assumes Δλ ≪ λ, ensuring high temporal coherence and thus well-defined, sharp intensity patterns without significant broadening from spectral dispersion. Without this condition, the finite imposes an envelope on the interference term, limiting the observable contrast to path differences on the of l_c; for instance, in white diffraction (Δλ ≈ λ), fringes are confined to separations of roughly λ² / Δλ ≈ λ, resulting in colorful but rapidly decaying patterns visible only near zero path difference. Temporal coherence is typically measured using a , where the visibility of fringes as a of path delay τ provides the of the field, with the time extracted from the width of the central peak. This setup reveals how polychromatic sources produce short-lived interference, contrasting with monochromatic ones./05:_Interference_and_coherence/5.05:_Temporal_Coherence_and_the_Michelson_Interferometer) Lasers exhibit high temporal due to their narrow linewidth (Δν ≪ 1/τ_c for long τ_c, often meters to kilometers), enabling over extended paths and precise diffraction measurements, whereas thermal sources like incandescent lamps have low (l_c on the order of micrometers) from broad blackbody spectra, restricting their use to short-baseline setups.

Applications

Optical Spectroscopy

Optical spectroscopy employs diffraction gratings to disperse light into its constituent wavelengths, enabling the analysis of spectral features such as absorption and emission lines for identifying elements and molecules. In grating spectrometers, light passes through an entrance slit and is collimated before striking the grating, where diffraction separates wavelengths at different angles, allowing detection via a focal plane or detector array. This technique revolutionized wavelength analysis by providing higher precision and broader spectral coverage compared to prisms. Historically, Joseph von Fraunhofer identified the dark absorption lines in the solar spectrum, now known as Fraunhofer lines, in 1814 using an early spectrometer, laying the foundation for diffraction-based spectral studies. The angular separation of wavelengths in a arises from the grating equation, leading to the d\theta = \frac{m}{d} \frac{d\lambda}{\cos\theta}, where m is the , d is the groove spacing, d\lambda is the wavelength difference, and \theta is the . This formula quantifies how small changes in wavelength produce measurable angular shifts, with higher orders m and finer groove spacing $1/d enhancing separation for closely spaced spectral features. In practice, the is blazed to maximize in a specific order, directing most into the desired diffraction direction. Resolution in grating spectrometers, defined as the smallest resolvable wavelength difference \Delta\lambda, is fundamentally limited by the slit width and groove density. Narrower entrance and exit slits reduce the effective bandwidth but increase by minimizing overlap between adjacent orders or lines, while higher groove densities (e.g., 1200 grooves/mm) yield finer and thus better \Delta\lambda, often achieving resolutions below 0.1 in optimized systems. The theoretical R = \lambda / \Delta\lambda approaches mN, where N is the total number of grooves illuminated, but practical limits arise from slit-induced broadening and grating imperfections. Echelle gratings extend this capability by operating in very high orders (typically m > 50), using coarse groove spacings (e.g., 30-80 grooves/mm) to achieve high resolving powers over wide ranges without requiring excessively long focal lengths. Blazed at steep angles near 63-76 degrees, echelles produce overlapping orders that are separated by a secondary element, such as a cross-disperser, enabling compact, broadband spectrometers with R > 10^5 for astronomical and laboratory applications. This design trades for resolution, making it ideal for detailed line profiling in complex spectra. Modern grating spectrometers often utilize the Czerny-Turner mount, a configuration featuring two spherical mirrors for collimation and focusing, which minimizes aberrations and provides a flat focal field for array detectors. Introduced in , this setup supports focal lengths from 0.2 to 1 m, balancing and throughput, and is prevalent in commercial instruments for UV-Vis-NIR due to its simplicity and correction.

Structural Biology

In structural biology, X-ray diffraction plays a pivotal role in elucidating the three-dimensional structures of biomolecules, particularly proteins, enabling insights into their function, interactions, and mechanisms of disease. , a cornerstone technique, involves growing high-quality crystals of purified proteins, which diffract s to produce patterns that reveal atomic arrangements. This method has been instrumental in determining nearly 250,000 protein structures deposited in the as of November 2025, with diffraction data underpinning the majority. The workflow of protein crystallography begins with protein expression and purification, followed by , where proteins are induced to form ordered lattices under controlled conditions such as varying pH, temperature, or precipitants. then occurs by exposing crystals to X-rays, typically at facilities that provide intense, tunable beams for high-resolution diffraction patterns; these sources have revolutionized the field by enabling rapid from microcrystals and reducing exposure times. Subsequent steps include indexing the diffraction spots, determining phases to reconstruct the map, and building an atomic model that is refined against the data to minimize discrepancies. radiation's brightness and tunability are essential, as they allow for wavelengths optimized to exploit anomalous scattering from atoms like for phasing. Despite these advances, challenges persist, notably the phase problem, where diffraction experiments measure only intensities (amplitudes squared) but not the phases of scattered waves, requiring indirect methods like isomorphous replacement or anomalous dispersion to reconstruct the structure. further complicates efforts, as absorption generates reactive species that disrupt bonds, decarboxylate side chains, and degrade over time, particularly in sensitive biomolecules; cryogenic cooling mitigates this but cannot eliminate it entirely. Seminal achievements underscore diffraction's impact: in 1958, John Kendrew reported the first three-dimensional , that of at 6 resolution, revealing its globular fold and pocket. This was followed in the 1960s by Max Perutz's determination of hemoglobin's structure at 5.5 , disclosing its quaternary arrangement and allosteric changes, which earned them the 1962 . In the 2020s, tools like have complemented diffraction by providing predictive models for molecular replacement phasing, accelerating structure solution for novel targets and validating experimental maps, though they do not replace empirical data. Beyond traditional , integration with (cryo-EM) has expanded capabilities through single-particle analysis, which reconstructs structures from thousands of flash-frozen molecules without crystals, offering complementary resolution for flexible or heterogeneous complexes. This hybrid approach leverages diffraction for high-resolution rigid domains while cryo-EM handles dynamic regions, as demonstrated in studies of large assemblies like ribosomes.

Modern Imaging Techniques

Modern imaging techniques in diffraction leverage advanced sources and computational methods to push beyond traditional limitations, enabling the study of dynamic and non-crystalline samples at unprecedented spatiotemporal resolutions. free-electron lasers (XFELs), such as the Linac Coherent Light Source (LCLS) operational since the 2010s, produce femtosecond-duration pulses that facilitate "diffraction before destruction" for radiation-sensitive, non-crystalline specimens. This approach captures diffraction patterns from individual particles or biomolecules in a single pulse before occurs, allowing serial femtosecond crystallography of structures that were previously inaccessible. Coherent diffraction imaging (CDI) reconstructs high-resolution images from measured intensity patterns using phase retrieval algorithms, which iteratively recover the lost phase information through constraints in real and reciprocal space. A key extension, , overlaps multiple coherent illuminations to enable robust three-dimensional reconstructions, achieving resolutions down to a few nanometers for extended samples like or biological assemblies. Techniques that surpass the classical diffraction limit, such as depletion (STED) microscopy and photoactivated localization microscopy (), exploit controlled coherence in illumination to confine excitation or localize emitters, yielding sub-100 nm resolutions in live-cell . In the 2020s, X-ray pulses have emerged for probing ultrafast dynamics, with recent demonstrations of atomic X-ray lasers generating 60-100 pulses to image quantum-scale phenomena without significant ionization damage. Complementing this, ultrafast (UED) advances, including high-repetition-rate sources and vortex beams, resolve in materials on to timescales at atomic resolutions. These methods offer advantages over traditional diffraction imaging, including minimal sample destruction via ultrashort pulses and enhanced resolutions approaching 1 nm, particularly for dynamic processes in non-crystalline environments.

References

  1. [1]
    27.2 Huygens's Principle: Diffraction – College Physics
    The bending of a wave around the edges of an opening or an obstacle is called diffraction. Diffraction is a wave characteristic and occurs for all types of ...
  2. [2]
    Diffraction of Sound - HyperPhysics
    Diffraction: the bending of waves around small obstacles and the spreading out of waves beyond small openings.
  3. [3]
    huygens_principle.html - UNLV Physics
    Christiaan Huygens (1629--1695) in 1678 proposed Huygens principle to explain the reflection and refraction of light. Augustin-Jean Fresnel (1788-1827) extended ...
  4. [4]
    A Brief History of Diffraction - Physics
    A Brief History of Diffraction. Diffraction describes the bending of waves around obstacles, or the bending of waves after passing through openings.
  5. [5]
    [PDF] Lecture 19: Diffraction and resolution
    Diffraction refers to what happens to a wave when it hits an obstacle. The key to understanding diffraction is a very simple observation first due to ...
  6. [6]
    The Feynman Lectures on Physics Vol. I Ch. 30: Diffraction - Caltech
    We shall now discuss the situation where there are n equally spaced oscillators, all of equal amplitude but different from one another in phase.
  7. [7]
    X-Ray Crystallography- History - Penn Arts & Sciences
    Max von Laue (1879-1960) German physicist won Nobel Prize in Physics in 1914 for his discovery of the diffraction of X-rays by crystals.
  8. [8]
    Wave Behaviors - NASA Science
    Aug 3, 2023 · An instrument called a spectrometer uses diffraction to separate light into a range of wavelengths—a spectrum. In the case of visible light, the ...<|separator|>
  9. [9]
    Francesco Maria Grimaldi - Linda Hall Library
    Apr 2, 2019 · Diffraction was first announced in Grimaldi's posthumous Physico-mathesis de lumine (1665), which we have in our History of Science Collection ( ...<|control11|><|separator|>
  10. [10]
    James Gregory - Biography - MacTutor - University of St Andrews
    The feather of a sea bird was to allow Gregory to make another fundamentally important scientific discovery while he worked in St Andrews. The feather ...
  11. [11]
    Rainbows, Halos, Coronas and Glories - AMS Journals
    The presence of halos indicates that polyhedral solid particles are present in the atmosphere, namely ice crystals. Any pair of crystal faces may act as a ...
  12. [12]
    Thomas Young and the Nature of Light - American Physical Society
    In May of 1801, while pondering some of Newton's experiments, Young came up with the basic idea for the now-famous double-slit experiment to demonstrate the ...
  13. [13]
    The Classical Double Slit Interference Experiment - Semantic Scholar
    The double slit experiment was first conceived of by the English physician-physicist Thomas Young in 1801. It was the first demonstrative proof that light ...
  14. [14]
    July 1816: Fresnel's Evidence for the Wave Theory of Light
    Jul 1, 2016 · The wave theory was making a comeback, thanks in part to the work of a French civil engineer named Augustin-Jean Fresnel.
  15. [15]
    Poisson's Spot (aka Spot of Arago)
    In 1818, Augustin Fresnel submitted a paper on the theory of diffraction for a competition sponsored by the French Academy. His theory ...Missing: Jean memoir original
  16. [16]
    Joseph von Fraunhofer (1787–1826) | High Altitude Observatory
    In 1821 Fraunhofer built the first diffraction grating, comprised of 260 close parallel wires. Well versed in the mathematical wave theory of light ...Missing: original | Show results with:original
  17. [17]
    The life and work of Joseph Fraunhofer (1787-1826) - AIP Publishing
    He ruled concentric circular lines on flat glass at high density, and reported seeing the dark lines as circular arcs with this grating. David Rittenhouse (an ...Missing: 1810s | Show results with:1810s
  18. [18]
    [PDF] XXXV. A Tentative Theory of Light Quanta. By LOUIS DE BROGLIE
    I shall in the present paper assume the real existence of light quanta, and try to see how it would be possible to reconcile with it the strong experimental ...
  19. [19]
    [PDF] Louis de Broglie - Nobel Lecture
    Thus to describe the properties of matter as well as those of light, waves and corpuscles have to be referred to at one and the same time. The electron can ...
  20. [20]
    [PDF] The discovery of the diffraction of X-rays by crystals
    '18 In t953 LAUE drew the logical inference, asserting that although many researchers had irradiated crystals with X-rays before FRIEDRICH and KNIPPING, they ...
  21. [21]
    The Project Gutenberg eBook of Treatise on Light, by Christiaan ...
    Treatise on Light in which are explained the causes of that which occurs in reflexion, & in refraction and particularly in the strange refraction of Iceland ...
  22. [22]
  23. [23]
    [PDF] Chapter 14 Interference and Diffraction - MIT
    The pattern is called a diffraction pattern. On the other hand, if no bending occurs and the light wave continue to travel in straight lines, then no ...
  24. [24]
    [PDF] Wave Optics 1 - Duke Physics
    Interference and diffraction of waves. The essential characteristic of energy transport by waves is that waves obey the superposition principle.
  25. [25]
    [PDF] Diffraction & Interference
    The phase difference δ (in radians) equals 2π x (extra distance)/λ where “extra distance” is the additional path traveled by light emanating from the slit ...
  26. [26]
    [PDF] PHYSICS 289 Experiment 9 Fall 2004 Wave Optics: Diffraction and ...
    All waves. (including light waves) are combined according to the superposition principle. This states that waves are combined by adding their amplitudes, not ...
  27. [27]
    [PDF] Mechanical waves - Duke Physics
    Write the phase difference in terms of the frequency: δ = (2π /λ)⋅Δx = (2π f /v)⋅Δx . This shows that δ increases as the frequency is raised. Zero ...
  28. [28]
    [PDF] List of useful formulae for Phys 123, Midterm 2 Density = Mass ...
    Phase difference due to path difference: δ = 2π ( Δx / λ ). For standing waves on a string fixed at both ends: fn = nf1 with f1 = v/2L and n = 1,2,3, … For ...
  29. [29]
    [PDF] Huygens principle; young interferometer; Fresnel diffraction
    The overall transmitted wavefront is the superposition of the. Huygens point sources obtained by point-by-point multiplication of the incident wavefront ...
  30. [30]
    [PDF] 17 Wave-Particle Duality
    Diffraction and interference phenomena reveal a wave- like character in light, but the photoelectric effect shows light to have a parti- clelike character as ...
  31. [31]
    [PDF] Physical Optics and Diffraction - Princeton University
    Fraunhofer regime (far field)​​ Neglecting the quadratic term: The amplitude of the Fraunhofer diffraction. pattern is given by the 2D Fourier transform of. the ...Missing: formulation | Show results with:formulation
  32. [32]
    Fourier Optics - UCSB Physics
    May 2, 2025 · It is equivalent to the more familiar Huygens-Fresnel principle which states that the diffracted field in the output plane ( x ′ y ′ x'y' x′y′- ...
  33. [33]
    [PDF] Fraunhofer Diffraction - UC Davis Math
    Formula (4.8) is obtained by considering the straight ray RA,2 parallel to (¤, », 1) t} passes through the point A, on the lens that lies on the optic axis.
  34. [34]
    Lecture 17: Fraunhofer diffraction; Fourier transforms and theorems
    Topics: Fraunhofer diffraction; review of Fourier transforms and theorems. Instructors: George Barbastathis, Colin Sheppard<|control11|><|separator|>
  35. [35]
    [PDF] 20. Diffraction and the Fourier Transform
    Fraunhofer Diffraction from a slit is simply the Fourier Transform of a rect function, which is a sinc function. The irradiance is then sinc2 . Page 15 ...Missing: definition | Show results with:definition
  36. [36]
  37. [37]
    [PDF] Introduction to Modern Optics - classe
    Definition of the variables for Fraunhofer diffraction by a single slit. and ... It follows that the fundamental diffraction integral [Equation (5.16)].
  38. [38]
    [PDF] 1 Huygens Integral
    An “obliquity factor K[θ(r, r0)]”has been added to the integral. It has to ... which is the 2 D extension of the expression (11). 4. Page 5. 3 Fraunhofer ...
  39. [39]
    [PDF] Chapter 8: Diffraction [version 1008.1.K] - Caltech PMA
    ... Fresnel half-period zones, of radius. √nrF , where n = 1, 2, 3 .... The ... observation point P on the optic axis, alternate half-period zones are obscured.<|control11|><|separator|>
  40. [40]
    Section 4.4 ZONE PLATES - X-Ray Data Booklet
    A zone plate is a circular diffraction grating. In its simplest form, a transmission Fresnel zone plate lens consists of alternate transparent and opaque rings.
  41. [41]
    Fraunhofer Single Slit Diffraction - HyperPhysics Concepts
    An element at one edge of the slit and one just past the centerline are chosen, and the condition for minimum light intensity is that light from these two ...
  42. [42]
    Single Slit Diffraction Intensity - HyperPhysics Concepts
    Under the Fraunhofer conditions, the wave arrives at the single slit as a plane wave. Divided into segments, each of which can be regarded as a point source, ...
  43. [43]
    Single slit diffraction
    Dark fringes in the diffraction pattern of a single slit are found at angles θ for which w sinθ = mλ, where m is an integer, m = 1, 2, 3, ... . For the first ...
  44. [44]
    Single-Slit Diffraction with a Laser | Physics III: Vibrations and Waves
    Apr 18, 2018 · 5:39with this single-slit experiment. 5:43And we call that diffraction. 5:47So you may be wondering, why do I call it diffraction? 5:50Why not interference ...
  45. [45]
    [PDF] interference and diffraction experiment
    Set up the laser at one end of the optical bench. Place the single slit disk in its holder about 3 cm in front of the laser and select the variable width slit.
  46. [46]
    Diffraction Gratings Tutorial - Thorlabs
    Sep 1, 2022 · Diffraction gratings, either transmissive or reflective, can separate different wavelengths of light using a repetitive structure embedded within the grating.Missing: Fraunhofer | Show results with:Fraunhofer
  47. [47]
    [PDF] Fraunhofer Diffraction at Diffraction Gratings
    Sep 14, 2015 · Figure 1: Diffraction in a grating. Diffraction is a phenomenon that is observed when light waves are obstructed by an obstacle such as a slit.
  48. [48]
    None
    Below is a merged summary of the key sections on diffraction gratings from the provided segments, combining all information into a concise yet comprehensive response. To retain maximum detail, I’ll use a table in CSV format for key formulas and concepts, followed by a narrative summary that integrates all additional details. This approach ensures all information is preserved while maintaining clarity and density.
  49. [49]
    Microscope Resolution: Concepts, Factors and Calculation
    Airy, G.B., On the Diffraction of an Object-Glass with Circular Aperture, Transactions Cambridge Philosophical Society (1835) vol. 5, part 3, pp. 283-291 ...<|control11|><|separator|>
  50. [50]
    Diffraction patterns of simple apertures - Optica Publishing Group
    It is well known that the Fraunhofer diffraction pattern of an aperture is given by the two-dimensional Fourier transform of that aperture.
  51. [51]
    Generalized Aperture and the Three-Dimensional Diffraction Image
    It is shown that when a lens produces an image of a point source the three-dimensional diffraction pattern which results is the three-dimensional Fourier ...
  52. [52]
    Fraunhofer diffraction at straight-edged apertures
    The two-dimensional Fourier transform of an arbitrary straight-edged aperture may be computed exactly from a simple expression in closed form, Eq. (4) . The ...
  53. [53]
    On Babinet's Principle and Diffraction Associated with an Arbitrary ...
    Babinet's principle states that, if the incident light propagates along the z-direction, the diffraction by a particle, shown in panel (a), can be replaced by ...
  54. [54]
    Efficient high-order accurate Fresnel diffraction via areal quadrature ...
    Oct 12, 2020 · It uses a high-order areal quadrature over the aperture, then exploits a single 2D nonuniform fast Fourier transform (NUFFT) to evaluate rapidly ...<|control11|><|separator|>
  55. [55]
    [PDF] Diffraction by Circular and Triangular Apertures as a Diagnostic Tool ...
    Oct 1, 2025 · We consider diffraction of twisted wave packets by a perfectly absorbing screen in the plane z = 0 with an open aperture area S of arbitrary ...
  56. [56]
    Diffraction of transmission light through triangular apertures in array ...
    However, an obtuse triangle aperture generates a deformed (or asymmetric) diffraction pattern, as shown in Fig. 7(c′) . There is one horizontal ghost line in ...
  57. [57]
    Analytic diffraction analysis of a 32-m telescope with hexagonal ...
    The Fourier transform of virtually any pupil function with only straight edges can be calculated by geometric integration and the properties of the Fourier ...
  58. [58]
    Convolution relation within the three-dimensional diffraction image
    The amplitude in the image of a point source is (a constant times) the three-dimensional Fourier transform of the lens pupil, i.e., of the amplitude ...
  59. [59]
    Gaussian beams - RP Photonics
    Kogelnik and T. Li, “Laser beams and resonators”, Appl. Opt. 5 (10), 1550 ... A. E. Siegman, Lasers, University Science Books, Mill Valley, CA (1986).
  60. [60]
    Press release: The Nobel Prize in Chemistry 2014 - NobelPrize.org
    Oct 8, 2014 · In 1873, the microscopist Ernst Abbe stipulated a physical limit for the maximum ... Image – Abbe's diffraction limit (pdf 447 kB) · Image ...
  61. [61]
    The Diffraction Barrier in Optical Microscopy | Nikon's MicroscopyU
    In terms of resolution, the radius of the diffraction Airy disk in the lateral (x,y) image plane is defined by the following formula: Formula 1 - Radius of ...
  62. [62]
    Microscopy Basics | The Point Spread Function - Zeiss Campus
    The ideal point spread function (PSF) is the three-dimensional diffraction pattern of light emitted from an infinitely small point source in the specimen.
  63. [63]
    Diffraction limit - Scientific Volume Imaging
    Diffraction limit. The resolution of optical microscopy is physically limited. This fundamental limit was first described by Ernst Karl Abbe in 1873 (1) and ...
  64. [64]
    The Airy Disk and Diffraction Limit
    ### Summary of Airy Disk, Resolution Limit, and Diffraction-Limited Imaging
  65. [65]
    Strehl ratio - Amateur Telescope Optics
    The Strehl ratio is the ratio of peak diffraction intensities of an aberrated vs. perfect wavefront, expressing the effect of wavefront aberrations on image ...
  66. [66]
    The Effect of a Parabolic Apodizer on Improving the Imaging ... - MDPI
    Dec 25, 2023 · We present the results of improving resolution in the imaging of two closely spaced point sources with an optical system under the influence of apodization.
  67. [67]
    Speckle Patterns - an overview | ScienceDirect Topics
    A speckle pattern is defined as a grainy light distribution that results from the self-interference of numerous waves scattered from a diffuse object's surface, ...
  68. [68]
    (PDF) Effect of Phase Relations on Speckle Pattern: Simulation and ...
    ... Speckle is derived from the coherent superposition of random wavelets. generated by a rough surface from a light source as shown in Fig. 1. Fig. 1 Speckle ...
  69. [69]
    Some fundamental properties of speckle* - Optica Publishing Group
    Goodman, “Statistical properties of laser speckle patterns,” in Laser speckle and related phenomena, edited by J. C. Dainty (Springer-Verlag, Heidelberg ...
  70. [70]
    Statistical Properties of Speckle Patterns - SpringerLink
    The purpose of this chapter is to introduce the various basic statistical properties of speckle patterns. We will basically follow Goodman [22] with some ...Missing: diffraction formation
  71. [71]
    First-Order Statistical Properties of Optical Speckle
    Contrast is a measure of how strong the fluctuations of intensity are in a speckle pattern compared with the average intensity. Signal-to-noise ratio is the ...
  72. [72]
    Speckle Pattern - Introduction and Applications - AZoOptics
    May 15, 2014 · Objective speckle pattern - Objective speckle patterns are formed when a diffuse object is illuminated by a coherent wave. The speckle pattern ...
  73. [73]
    Surface-roughness study using laser speckle method - ScienceDirect
    In this method, two speckle patterns are recorded on the same film by varying the angle of illumination of an unexpanded laser beam on the surface. Placing the ...
  74. [74]
    Surface-roughness Measurement Based on the Intensity Correlation ...
    The statistical properties of speckle patterns generated from a rough surface under a fully developed static speckle-pattern illumination are examined.
  75. [75]
    Measurement of the roughness surface using the normalized ...
    We present a new method of measure of the roughness based on the analysis of the texture of speckle pattern on the surface. Images of speckle pattern over ...
  76. [76]
    [PDF] Babinet's Principle for Electromagnetic Fields 1 Problem 2 Solution
    In optics, Babinet's principle [1] for complementary screens is that the sum of the wave transmitted through a screen (usually considered to be “black” ...
  77. [77]
  78. [78]
    [PDF] Diffraction
    Roughly sketch the irradiance distribution for the Fresnel diffraction pattern of a knife edge. Be sure to point out the edge of the geometrical shadow.
  79. [79]
    Cornu's spiral in the Fresnel regime studied using ultrasound
    Mar 15, 2012 · A. Graphical treatment of diffraction: The Cornu spiral. Consider diffraction of a plane wave by a semi-infinite plane obstacle (a half plane) ...Missing: primary | Show results with:primary
  80. [80]
    [PDF] Geometrical theory of diffraction - Indian Academy of Sciences
    This has been done specifically by Keller (1962) who exploited the Sommerfeld solution of diffraction of electromagnetic waves at a half plane, making the.
  81. [81]
    [PDF] Propagation by diffraction - ITU
    Step 1: Calculate the geometrical parameter ν for each of the three knife-edges (top, left side and right side) using any of equations (26) to (29). Step 2: ...
  82. [82]
    [PDF] Chapter 7 Diffraction - Caltech PMA
    7.3, we shall illustrate Fraunhofer diffraction by computing the expected angular resolution of the. Hubble Space Telescope, and in Sec. 7.4, we shall analyze ...
  83. [83]
    Calculation of vectorial diffraction in optical systems
    Vectorial diffraction theory is fundamental to the study of optical systems that have a high numerical aperture (NA) or use specific polarization, such as ...
  84. [84]
    [PDF] Numerical techniques for Fresnel diffraction in computational ...
    The Nyquist sampling theorem states that the required minimum sampling rate for the hologram should be twice that of the spatial frequency [BH93], thus.
  85. [85]
    Sample Pages (PDF) - SPIE Digital Library
    Our goal here is to illustrate the use of the convolution method for calculating diffraction patterns. Let the aperture of interest be square with width ℓ on a ...
  86. [86]
    [PDF] The Duality of Light - Wooster Physics
    May 7, 2020 · Qualitative Analysis of Intensity. The single slit intensity pattern is given by a sinc(x) function. We can plot this and see a large central ...
  87. [87]
    Fraunhofer Single Slit Diffraction
    ### Summary of Single Slit Diffraction Angle Formula
  88. [88]
    [PDF] Physics of Light and Optics
    This book covers light propagation, reflection, polarization, dispersion, coherence, ray optics, imaging, diffraction, and the quantum nature of light.
  89. [89]
  90. [90]
    Near Field and Far Field - RP Photonics
    Near field is the optical field close to an object or beam focus; far field is at large distances, with context-dependent meanings.
  91. [91]
    12.6 Limits of Resolution: The Rayleigh Criterion
    Any beam of light having a finite diameter D and a wavelength λ exhibits diffraction spreading. The beam spreads out with an angle θ given by the equation θ = ...
  92. [92]
    DeBroglie Wavelength - HyperPhysics
    For an electron with KE = 1 eV and rest mass energy 0.511 MeV, the associated DeBroglie wavelength is 1.23 nm, about a thousand times smaller than a 1 eV photon ...
  93. [93]
    [PDF] Electron Diffraction Using Transmission Electron Microscopy
    Electron diffraction via the transmission electron microscope is a powerful method for characterizing the structure of materials, including perfect crystals ...
  94. [94]
    Diffraction of Electrons by a Crystal of Nickel | Phys. Rev.
    Feb 3, 2025 · Davisson and Germer showed that electrons scatter from a crystal the way x rays do, proving that particles of matter can act like waves. See ...
  95. [95]
    [PDF] Introduction to Neutron Scattering
    Neutrons are used because they have short-range nuclear interactions, are penetrating, don't destroy samples, and their wavelengths are comparable to atomic ...
  96. [96]
    Neutron temperature - Wikipedia
    Neutron energy distribution ranges ; 25 meV, ≈ 1.8 Å, Thermal neutrons (at 20 °C) ; 25 meV–0.4 eV, ≈ 1.8 - 0.45 Å, Epithermal neutrons ; 10–300 eV, ≈ 0.09 - 0.016 ...<|control11|><|separator|>
  97. [97]
    The early development of neutron diffraction: science in the wings of ...
    Ernest O. Wollan, Lyle B. Borst and Walter H. Zinn were all able to observe neutron diffraction in 1944 using the X-10 graphite reactor and the CP ...Missing: 1940s | Show results with:1940s
  98. [98]
    A History of Neutron Scattering at ORNL
    Jun 13, 2018 · Neutron scattering grew from the nuclear science of the Manhattan Project during the 1940s at what is now Oak Ridge National Laboratory (ORNL).
  99. [99]
    Shull and Wollan, neutron pioneers - ORNL and the Nobel Prize
    Clifford Shull (left) and Ernest Wollan (right) conducted some of the world's first neutron scattering experiments using a diffractometer, installed at ORNL's ...Missing: history | Show results with:history
  100. [100]
    What makes neutron scattering unique | ORNL
    Apr 1, 2020 · Perhaps an even more important upside to this weak scattering is that neutrons easily penetrate materials that would stop X-rays or electrons, ...
  101. [101]
    [PDF] Neutron Scattering - A Primer
    Reactor Sources. Neutrons are pro- duced in a nuclear reactor by the fis- sioning of atoms in the reactor fuel, which, for research reactors, is invari-.
  102. [102]
    Magnetic structures description and determination by neutron ...
    Neutron diffraction determines magnetic structures by finding where moments point, using neutron's mass and electrical neutrality to probe the sample.
  103. [103]
    [PDF] Neutron powder diffraction
    ... Debye-Scherrer" cones of scattering at angles 29 such that sin 9 = / 2dhkI where the plane spacingd hk ^. =2n I |G hk J . 1.4. Scattering of Neutronsand X-rays.
  104. [104]
    Current status of neutron crystallography in structural biology - PMC
    Neutron crystallography enables direct observation of hydrogen atoms, and reveals molecular recognition through hydrogen bonding and catalytic reactions ...Diffraction Data Measurement... · Figure 1 · Figure 3
  105. [105]
    Enhanced visibility of hydrogen atoms by neutron crystallography on ...
    Neutron crystallography is capable of directly revealing the position of hydrogens, but its use on unlabeled samples faces certain technical difficulties: the ...Materials And Methods · Crystallization And... · Results
  106. [106]
    [PDF] WILLIAM LAWRENCE BRAGG The Dawn of X-ray Crystallography
    Here we see for the first time the appearance of Bragg's Law. In its initial form he chooses the angle θ to be the angle of incidence which gives the cosine ...Missing: derivation | Show results with:derivation
  107. [107]
    Max von Laue and the discovery of X‐ray diffraction in 1912 - 2012
    First published: 07 May 2012. https://doi.org/10.1002/andp.201200724. Citations: 48. About. References. Related. Information. PDF · PDF. Tools.
  108. [108]
    Zur Begründung der Kristalloptik - Ewald - 1916 - Wiley Online Library
    Zur Begründung der Kristalloptik. P. P. Ewald,. P. P. Ewald. Search for more ... Download PDF. back. Additional links. About Wiley Online Library. Privacy ...
  109. [109]
    [PDF] The diffraction of X-rays by crystals - Nobel Lecture, September 6 ...
    As a pulse passes over each diffracting point, it scatters a wave, and if a number of points are arranged on a plane the diffract- ed wavelets will combine ...Missing: 1913 | Show results with:1913
  110. [110]
    The reflection of X-rays by crystals - Journals
    The reflection of X-rays by crystals. William Henry Bragg. Google Scholar · Find ... This Issue. 01 July 1913. Volume 88Issue 605. Article Information. DOI ...Missing: derivation | Show results with:derivation
  111. [111]
    1913: The First Steps | Early Days of X-ray Crystallography
    In the first part of his talk, M. von Laue presented the theory of X‐ray diffraction by crystals, including the three Laue conditions, Bragg's law, the ...
  112. [112]
    100 years of the X-ray powder diffraction method | OUPblog
    Sep 20, 2016 · In a series of papers published in 1913, Debye calculated the influence of lattice vibrations on the diffracted intensity (the Debye factor), ...
  113. [113]
    Introduction to phasing - PMC - NIH
    2.3. Isomorphous replacement. The use of heavy-atom substitution to solve the phase problem was invented very early on by small-molecule crystallographers, for ...
  114. [114]
    Degree of Coherence & Optical Problems - ScienceDirect
    The maximum visibility of the interferences obtainable from two points in a wave field is defined as their degree of coherence γ.Missing: Physics | Show results with:Physics
  115. [115]
    [PDF] Coherent X-ray Scattering and X-ray Photon Correlation Spectroscopy
    λ ξ σ π. = More exactly: for a Gaussian source profile of width σ the coherence length of the source is given by: At current third generation synchrotron ...<|control11|><|separator|>
  116. [116]
    Investigation of spatial coherence and fringe visibility - AIP Publishing
    Sep 1, 2018 · We discuss Young's double-slit experiment using a partially coherent light source consisting of a helium-neon laser incident on a rotating piece of white paper.INTRODUCTION · Negligible slit widths · Finite slit width · A more complete model
  117. [117]
    Chapter 10 - Imaging with Quasi-Monochromatic Waves - GlobalSpec
    imaging with quasi-monochromatic coherent or incoherent wave fields. Monochromatic wave fields have a single temporal frequency f. Nonmonochromatic wave fields ...
  118. [118]
    coherent, light, spatial and temporal coherence, monochromaticity
    Lasers typically exhibit very high spatial coherence, resulting in excellent beam quality. However, their temporal coherence varies: Single-frequency lasers ...
  119. [119]
    Joseph von Fraunhofer
    He was the first to measure the spectrum of sunlight and characterize the dark absorption strips it contains: the »Fraunhofer lines«. His independent, ...Missing: 1810s | Show results with:1810s
  120. [120]
    The Era of Classical Spectroscopy - MIT
    Fraunhofer also developed the diffraction grating, an array of slits, which disperses light in much the same way, as does a glass prism, but with important ...
  121. [121]
    Diffraction Grating Physics - Newport
    A diffraction grating is a multi-slit surface that provides angular dispersion, separating wavelengths based on the angle they emerge from the grating.Missing: dθ/ dλ = θ)
  122. [122]
    [PDF] Astronomical Techniques I - Lecture 12 - NCRA-TIFR
    dλ. = n d cosθ. Higher dispersion corresponds to larger values of this quantity. Echelles take advantage of both n and θ dependence to maximize dispersion.
  123. [123]
    Resolution of Monochromators and Spectrographs - Newport
    Higher groove densities give higher reciprocal dispersion and therefore higher resolution. The grating dispersion is similar for gratings with the same groove ...
  124. [124]
    Spectral Resolution and Dispersion in Raman Spectroscopy
    Sep 1, 2020 · The groove density of the grating in conjunction with the focal length of the spectrometer determines the dispersion (dλ/dl) or spread of the ...
  125. [125]
    Echelle Grating Spectroscopic Technology for High-Resolution and ...
    Echelle grating provides high spectral resolving power and diffraction efficiency in a broadband wavelength range by the Littrow mode.
  126. [126]
    Echelle Gratings - an overview | ScienceDirect Topics
    These gratings operate with relatively high orders (m ∼ 1000). A 10 in. echelle with 100 grooves per inch, designed for observation in the 1000th order ...
  127. [127]
    Monochromators - Shimadzu Scientific Instruments
    The Czerny-Turner mount uses two symmetrically arranged spherical mirrors as the collimating mirror and camera mirror, as shown in Fig. 10. A concave mount ...
  128. [128]
    Broadband astigmatism-free Czerny-Turner imaging spectrometer ...
    The Czerny–Turner spectrometer is a commonly used instrument for resolving the spectral intensity of radiation imaged across one spatial dimension. Modern ...
  129. [129]
    Synchrotron Radiation as a Tool for Macromolecular X-Ray ...
    Synchrotron beamlines have been used to determine over 70% of all macromolecular structures deposited into the Protein Data Bank (PDB). These structures were ...
  130. [130]
    Protein crystallography for aspiring crystallographers or how to ...
    The present review provides an outline of the technical aspects of crystallography ... AutoDrug: fully automated macromolecular crystallography workflows ...
  131. [131]
    A beginner's guide to radiation damage - PMC - PubMed Central
    Irradiated protein crystals suffer an overall loss of resolution as high-angle spots fade away which is referred to as global damage. There are also specific ...
  132. [132]
    A Three-Dimensional Model of the Myoglobin Molecule Obtained by ...
    In 1958, J. C. Kendrew et al. applied Perutz–s technique to produce the first three-dimensional images of any protein - myoglobin, the protein used by ...
  133. [133]
    Structure of haemoglobin: a three-dimensional Fourier ... - PubMed
    Structure of haemoglobin: a three-dimensional Fourier synthesis at 5.5-A. resolution, obtained by X-ray analysis. Nature. 1960 Feb 13;185(4711):416-22. doi: ...
  134. [134]
    AlphaFold2: Crystallographic Phasing by Molecular Replacement
    The implications of the AlphaFold2 protein structure-modelling software for crystallographic phasing strategies are discussed.
  135. [135]
    How cryo‐electron microscopy and X‐ray crystallography ... - NIH
    In this review we provide some examples, and discuss how X‐ray crystallography and cryo‐EM can be combined in deciphering structures of macromolecules.
  136. [136]
    Serial Femtosecond Crystallography
    Thanks to the diffraction before destruction principle, there is no need to mitigate the conventional X-ray radiation damage at LCLS. As a result ...
  137. [137]
    Linac Coherent Light Source: The first five years | Rev. Mod. Phys.
    Mar 9, 2016 · Today the diffraction-before-destruction method is widely used at LCLS, and a large and growing literature describes simulations of damage ...
  138. [138]
    Diffraction before destruction | Nature Methods
    Mar 30, 2011 · Diffraction before destruction ... Two recent reports demonstrate the first applications of X-ray free-electron lasers to look at biological ...Missing: facility | Show results with:facility
  139. [139]
    Coherent Diffractive Imaging (CDI) - UCLA Physics & Astronomy
    (F) Phase retrieval algorithms iterate back and forth between real and reciprocal space. In each iteration, various constraints, including support, positivity ( ...
  140. [140]
    Direct three-dimensional tomographic reconstruction and phase ...
    Feb 1, 2019 · Three-dimensional (3D) ptychography or ptychotomography extends the unique capabilities of CDI techniques to a higher spatial dimension, relying ...
  141. [141]
    Breaking the Diffraction Barrier: Super-Resolution Imaging of Cells
    In this Primer, we explain the principles of various super-resolution approaches, such as STED, (S)SIM, and STORM/(F)PALM. Then, we describe recent applications ...Missing: coherence | Show results with:coherence
  142. [142]
    Scientists create first attosecond atomic X-ray laser
    Jun 11, 2025 · This breakthrough creating an atomic X-ray laser with sub-100 attosecond pulses allows researchers to study how electrons move inside atoms with ...
  143. [143]
    Ultrafast electron diffraction: Visualizing dynamic states of matter
    Dec 6, 2022 · Recent developments in the generation, manipulation, and characterization of ultrashort electron beams aimed at improving the combined ...
  144. [144]
    Pushing the limits of ultrafast diffraction: Imaging quantum ...
    Sep 20, 2024 · This article discusses advances in ultrafast X-ray scattering and electron diffraction, highlighting their potential to resolve attosecond charge migration and ...
  145. [145]
    [PDF] Diffractive Imaging Using Coherent X-ray Light Sources
    First, by taking advantage of the extremely short XFEL and HHG pulses, CDI is ideally suitable to probe functioning systems at the nanoscale - at multiple sites ...
  146. [146]
    Coherent X-ray diffractive imaging: applications and limitations
    In material science, where samples are less sensitive to radiation damage, we expect CXDI to be able to achieve 1 to 2 nm resolution using modern x-ray ...
  147. [147]
    University Physics Volume 3: 4.5 Circular Apertures and Resolution
    OpenStax textbook section on diffraction through circular apertures, including the Airy pattern, intensity distribution, and resolution limits.