Fact-checked by Grok 2 weeks ago

Airy disk

The Airy disk, also known as the Airy disc or Airy pattern, is the characteristic diffraction pattern formed when a point source of monochromatic light passes through a circular aperture in an ideal, aberration-free optical system, resulting in a bright central spot surrounded by concentric rings of progressively fainter intensity. This pattern arises due to the wave nature of light and the interference effects at the aperture edges, representing the fundamental limit of optical resolution rather than a true point image. Named after the English astronomer and mathematician George Biddell Airy, who provided the first complete theoretical derivation of the pattern in his 1835 paper "On the Diffraction of an Object-glass with Circular Aperture," the Airy disk quantifies the diffraction limit in imaging systems such as microscopes, telescopes, and cameras. Airy's analysis, based on the Huygens-Fresnel principle and expressed through Bessel functions, showed that approximately 84% of the light energy is concentrated in the central disk, with the remainder distributed in the surrounding rings. The radius of the Airy disk to its first minimum (the boundary of the central spot) is given by r = 1.22 \lambda f/\#, where \lambda is the wavelength of light and f/\# is the f-number of the optical system; equivalently, in microscopy, it is r = 0.61 \lambda / \mathrm{NA}, with NA denoting the numerical aperture. The Airy disk plays a critical role in defining the diffraction limit, the theoretical maximum resolving power of an optical instrument, as two point sources are resolvable only if separated by at least the Rayleigh criterion distance of $1.22 \lambda f/\#, where their disks overlap such that the central maximum of one falls on the first minimum of the other, yielding about 26% contrast at the midpoint. This limit scales inversely with aperture size and wavelength, meaning shorter wavelengths (e.g., blue light at 400 nm) or larger apertures produce smaller disks and higher resolution, but it imposes an unavoidable blur on images of fine details. In practice, factors like aberrations, illumination type (e.g., Köhler vs. critical), and partial coherence can modify the effective Airy disk size, but the ideal pattern remains the benchmark for evaluating optical performance across fields including astronomy, where it affects planetary and stellar imaging, and microscopy, where it governs the ability to distinguish sub-micron structures.

Fundamentals

Definition and Origin

The Airy disk is the brightest central feature of the Airy pattern, which arises as the Fraunhofer diffraction pattern from a uniformly illuminated circular aperture under monochromatic plane wave illumination. This central spot, surrounded by concentric darker and brighter rings, forms the point spread function for an ideal optical system with a circular pupil. Physically, the Airy disk originates from the interference of diffracted wavefronts in the focal plane of a lens, where light passing through the aperture bends and overlaps due to its wave properties. According to the Huygens-Fresnel principle, each point across the aperture acts as a secondary source of spherical wavelets, and their coherent superposition leads to constructive interference at the center, creating the bright disk while destructive interference produces the surrounding pattern. This phenomenon underscores the wave nature of light and establishes the diffraction limit, which sets the scale of the smallest resolvable details for a given wavelength and aperture size in aberration-free systems. Diffraction patterns vary with aperture geometry; a single-slit (rectangular) aperture yields a linear central maximum flanked by symmetric minima and secondary maxima, whereas the circular aperture's rotational symmetry results in the radial Airy pattern with its prominent central disk. The Huygens-Fresnel principle provides the foundational mechanism for both, treating the aperture as a collection of point sources whose wavelets interfere to shape the far-field distribution. Observations of the Airy disk were integral to 19th-century investigations confirming light's wave behavior through diffraction experiments, laying the groundwork for understanding resolution constraints in microscopy and telescopy.

Historical Development

The concept of the Airy disk emerged from early 19th-century advancements in wave optics, building on foundational experiments that established light's wave nature. In 1801, Thomas Young demonstrated interference through his double-slit experiment, providing empirical support for the wave theory of light and laying groundwork for understanding diffraction patterns. This was further developed by Augustin-Jean Fresnel in 1818, whose mathematical treatment of diffraction using Huygens' principle explained phenomena like the bending of light around obstacles, setting the stage for precise modeling of aperture effects in optical systems. The Airy disk itself was first mathematically derived in 1835 by George Biddell Airy, then Astronomer Royal, in his paper "On the Diffraction of an Object-glass with Circular Aperture," published in the Transactions of the Cambridge Philosophical Society. Airy analyzed the diffraction pattern produced by a circular aperture in telescopes, describing the central bright spot surrounded by concentric rings as the fundamental limit of image resolution under wave optics. His work directly applied Fresnel's diffraction theory to practical astronomical instruments, highlighting how imperfections in focusing were not due to lens flaws but inherent wave propagation. In the 1870s, Lord Rayleigh (John William Strutt) provided experimental confirmation and theoretical refinement of Airy's pattern, linking it explicitly to optical resolution limits. In his 1879 series of papers "Investigations in Optics, with Special Reference to the Spectroscope" in the Philosophical Magazine, Rayleigh explored how the Airy disk determines the minimum separable distance between point sources in telescopes and spectroscopes, establishing the Rayleigh criterion where two disks are just resolvable when the central maximum of one coincides with the first minimum of the other. This criterion, developed through his analyses in the late 1870s and 1880s, profoundly influenced the design of optical instruments by quantifying diffraction's impact on resolving power. Rayleigh further extended these ideas to microscopy in his 1896 memoir "On the Theory of Optical Images, with Special Reference to the Microscope," confirming the Airy disk's role in microscopic resolution and reconciling it with Ernst Abbe's earlier work. Concurrently, computational modeling advanced its application, enabling numerical simulations of diffraction in complex systems from the 1960s, which facilitated optimizations in laser optics and high-resolution imaging without physical prototypes. These extensions solidified the Airy disk as a cornerstone for modern optical engineering.

Mathematical Description

Intensity Pattern Formulation

The intensity pattern of the Airy disk arises from the Fraunhofer diffraction of a plane wave through a circular aperture, which can be derived using the Fourier transform of the aperture's pupil function. The pupil function P(\xi, \eta) for a circular aperture of radius a (diameter D = 2a) is given by P(\xi, \eta) = \text{circ}\left(\sqrt{\xi^2 + \eta^2}/a\right), where \text{circ}(r) = 1 for r \leq 1 and 0 otherwise, assuming uniform illumination. In the Fraunhofer approximation, valid for far-field conditions (observation distance much larger than aperture size and wavelength), the complex amplitude U(x, y) in the observation plane is the Fourier transform of the pupil function: U(x, y) = \frac{1}{i \lambda f} \iint P(\xi, \eta) \exp\left[-i \frac{2\pi}{\lambda f} (x \xi + y \eta)\right] d\xi \, d\eta, where \lambda is the wavelength, f is the focal length (or effective distance), and spatial frequencies are u = x/(\lambda f), v = y/(\lambda f). Due to circular symmetry, this reduces to a Hankel transform (Fourier-Bessel transform) in polar coordinates (\rho, \phi) for the aperture and (r, \theta) for the observation plane: U(r) \propto \int_0^a \int_0^{2\pi} \rho \, d\rho \, d\phi \, \exp\left[-i \frac{2\pi r \rho}{\lambda f} \cos(\phi - \theta)\right] = 2\pi \int_0^a J_0\left(\frac{2\pi r \rho}{\lambda f}\right) \rho \, d\rho, where J_0 is the zeroth-order Bessel function of the first kind. The integral evaluates to U(r) \propto a^2 \frac{J_1(ka \sin\theta)}{ka \sin\theta}, with k = 2\pi / \lambda the wave number and \theta the angular radius from the optical axis (small-angle approximation \sin\theta \approx r/f). Normalizing by the on-axis value (\theta = 0) yields the amplitude U(\theta) = U_0 \frac{2 J_1(x)}{x}, where x = \frac{\pi D \sin\theta}{\lambda} is the normalized radial coordinate, and U_0 is the central amplitude. The intensity pattern, I(\theta) = |U(\theta)|^2, is then I(\theta) = I_0 \left[ \frac{2 J_1(x)}{x} \right]^2, where I_0 = |U_0|^2 is the central intensity. This Airy pattern features a bright central disk surrounded by alternating bright and dark concentric rings, with dark rings occurring at the zeros of J_1(x), the first at x \approx 3.832. The central disk, bounded by this first zero, contains approximately 83.8% of the total diffracted energy, as determined by integrating the intensity over the pattern (total energy \propto \pi I_0 a^2).

Size and Radius Calculation

The radius of the Airy disk is defined as the radial distance from the center of the diffraction pattern to its first intensity minimum, which occurs where the argument x = \frac{\pi D \sin \theta}{\lambda} = 3.8317, with D the diameter of the circular aperture, \lambda the wavelength of light, and \theta the angular displacement from the optical axis. For small angles, \sin \theta \approx \theta, yielding the angular radius \theta \approx 1.22 \lambda / D. This value arises from the first zero of the first-order Bessel function of the first kind, J_1(3.8317) = 0, which governs the radial intensity distribution in the diffraction pattern. In the focal plane of a lens with focal length f, the linear radius r of the Airy disk is obtained by projecting the angular radius onto the plane, giving r = f \theta \approx 1.22 \lambda f / D. The full width (diameter) of the central disk is thus approximately $2.44 \lambda f / D. These dimensions scale inversely with aperture diameter D, emphasizing the role of larger apertures in reducing the spot size, and directly with wavelength \lambda and focal length f. The central Airy disk encloses approximately 83.8% of the total diffracted energy, with the remaining energy distributed in the surrounding concentric rings. This concentration highlights the disk's dominance in the pattern, as the first bright ring captures only about 7.2% of the energy. The size of the Airy disk establishes the fundamental limit of optical resolution, as articulated in the Rayleigh criterion: two point sources are just resolvable when separated by an angular distance equal to the disk's radius, $1.22 \lambda / D, such that the central maximum of one pattern coincides with the first minimum of the other, producing a detectable intensity dip. This criterion, derived from the overlap of Airy patterns, implies that finer separation results in indistinguishable blurring. In practical terms, for visible light (\lambda \approx 0.55 \, \mum) and a microscope objective with numerical aperture NA \approx 1.4 (corresponding to an effective D/(2f) \approx 1.4), the linear radius scales to roughly 0.2 \mum, setting the micron-scale limit for resolving cellular structures.

Practical Applications and Examples

Imaging Systems like Cameras

In imaging systems such as cameras and telescopes, the Airy disk defines the fundamental diffraction limit, which establishes the finest detail that can be resolved regardless of lens quality or sensor capabilities. This limit arises because light diffracting through the circular aperture spreads into the characteristic central bright spot surrounded by faint rings, blurring point sources and reducing overall image sharpness. For photographic cameras, the diffraction limit influences the choice of aperture settings, as smaller apertures (higher f-numbers) increase the Airy disk size on the sensor plane, leading to noticeable softening. To maintain sharpness, photographers typically avoid exceeding a minimum f-number where the Airy disk diameter approaches or exceeds the circle of confusion or pixel pitch; for instance, in a 35 mm focal length lens with a 10 mm aperture diameter at a wavelength of 550 nm, the angular radius of the Airy disk is approximately 13.6 arcseconds, setting a practical bound for resolvable detail. Telescopic imaging systems, operating at much larger scales, achieve superior resolution due to their expansive apertures, but the Airy disk still imposes the ultimate constraint, often trading off against other aberrations like spherical or chromatic errors. The Hubble Space Telescope, with its 2.4 m primary mirror, attains a diffraction-limited angular resolution of about 0.05 arcseconds at visible wavelengths, enabling the capture of fine structures in distant galaxies and nebulae that ground-based telescopes cannot resolve without adaptive optics. This performance highlights the Airy disk's role in balancing light-gathering power with precision, where larger diameters shrink the disk size proportionally, though manufacturing imperfections or atmospheric effects (mitigated in space) can degrade the theoretical limit. In digital imaging, the Airy disk's size must align with sensor pixel dimensions to prevent undersampling, where the diffraction pattern spills across multiple pixels, reducing contrast and effective resolution. Optimal design matches the Airy disk radius to roughly the pixel pitch, ensuring each point source illuminates at least 2-3 pixels for accurate sampling; undersampling occurs if pixels are too large relative to the disk, blurring fine details, while oversampling with tiny pixels yields diminishing returns due to noise. For example, in high-resolution full-frame sensors with 4-6 μm pixels, diffraction limits sharpness beyond f/8-f/11, as the Airy disk expands to encompass multiple pixels. Smartphone cameras exemplify this challenge, where compact lenses and small sensors (often 1/2.3-inch or smaller) with pixel sizes around 1-2 μm make diffraction dominant at relatively high f-stops like f/8 or beyond, as the Airy disk quickly overtakes pixel spacing and limits megapixel advantages. In these systems, wide apertures (low f-numbers, e.g., f/1.8) are favored for low-light performance, but computational processing partially compensates for diffraction-induced blur in post-capture sharpening. To mitigate the Airy disk's ringing artifacts, which can introduce unwanted halos around bright points in images, techniques like apodization or pupil masking modify the aperture's transmission profile, concentrating more energy in the central disk while suppressing secondary rings. Apodization, often implemented via graded neutral-density filters over the pupil, can suppress the first bright ring intensity by more than a factor of 4 in some designs, improving contrast in high-dynamic-range scenes, though at the cost of slightly enlarging the central spot and reducing total light throughput. Pupil masking, used in advanced telescopic systems, creates apodized shapes to further suppress diffraction for exoplanet imaging.

Human Vision

The human eye operates as a diffraction-limited optical system, with the pupil serving as the circular aperture that produces an Airy disk pattern on the retina for point sources of light. The pupil diameter typically ranges from 2 to 8 mm, constricting to about 2-4 mm under bright photopic conditions and dilating to 6-8 mm in dim scotopic environments, directly influencing the size of the Airy disk. The angular size of the Airy disk sets the theoretical diffraction limit for resolution, approximately 20-70 arcseconds depending on pupil diameter and visible wavelengths around 550 nm, with smaller pupils in bright light yielding larger disks and greater blur for fine details. In practice, however, human visual acuity averages about 1 arcminute (60 arcseconds) for standard 20/20 vision, as retinal cone spacing—around 0.5 arcminutes in the fovea—and neural processing impose additional constraints beyond pure optical diffraction. Interestingly, certain aspects of perception surpass this diffraction limit through higher-level neural mechanisms; for instance, vernier acuity, which detects subtle misalignments between lines, can resolve offsets as small as 5-10 arcseconds by pooling positional signals from multiple cones in the visual cortex. This hyperacuity enables finer relative judgments than absolute resolution would suggest. In scotopic vision, the larger pupil reduces diffraction blur, but acuity remains limited by sparser rod photoreceptors, emphasizing the interplay between optical and biological factors. A practical manifestation of these limits is the eye's inability to resolve stars separated by less than about 1 arcminute, as their overlapping Airy disks appear as a single point of light, aligning with the 20/20 threshold where details subtending 1 arcminute become distinguishable. Thus, while the Airy disk defines the fundamental optical boundary, human vision's effective performance reflects a balance of diffraction, retinal sampling, and cortical integration.

Laser Beams and Optical Devices

In applications involving focused laser beams, a collimated beam passing through a circular aperture and then focused by a lens forms an Airy pattern in the focal plane, with the central disk representing the diffraction-limited spot size achievable by the system. This spot size sets a fundamental limit on the precision of beam concentration, as even aberration-free optics cannot produce a smaller focus due to wave diffraction. The high spatial coherence of laser light further enhances the visibility and sharpness of the Airy pattern, allowing the concentric rings to be distinctly observed without the blurring from incoherent superposition. In practical devices such as laser cutting and welding systems, this diffraction-limited spot size typically ranges from tens of microns, depending on the laser wavelength and focusing optics; for instance, ytterbium fiber lasers at 1070 nm achieve spots around 50-100 μm to balance power density with material processing efficiency. Smaller spots enable finer cuts but are constrained by the Airy disk, which increases with longer wavelengths or lower numerical apertures. For aiming sights in rifle scopes and laser pointers, the Airy pattern manifests as beam divergence and visual blooming from the diffraction rings, where the surrounding intensity lobes create a halo effect, particularly noticeable in low-light conditions or on diffuse surfaces. A typical green laser pointer operating at 532 nm with a 1 mm aperture exhibits a diffraction-limited divergence of approximately 0.65 mrad (calculated as 1.22 λ / D), though practical designs often yield around 1 mrad due to beam truncation and Gaussian profiles. The Airy disk also plays a critical role in precision applications like optical trapping, where the spot size limits the confinement volume for microscopic particles or atoms; for a 1064 nm beam with numerical aperture 0.43, the Airy radius of approximately 1.51 μm (given by 0.61 λ / NA) determines the trap's transverse resolution and stiffness. Similarly, in LIDAR systems, the Airy disk governs angular resolution, as the diffraction pattern's size at the receiver affects the ability to distinguish closely spaced targets, with single-photon implementations matching pixel sizes to the Airy diameter for optimal performance.

Observation and Visibility

Conditions for Clear Observation

To observe the Airy disk distinctly in experimental settings, monochromatic illumination is essential, as it produces a sharp central disk surrounded by well-defined concentric rings; polychromatic light causes blurring of these rings due to the spread in wavelengths, resulting in overlapping and partially canceling diffraction patterns. A helium-neon laser, emitting at 632.8 nm, is commonly used for this purpose because it provides highly coherent, monochromatic light suitable for generating clear patterns. Plane wave incidence, achieved with a collimated beam, is required to produce the ideal Fraunhofer diffraction pattern characteristic of the Airy disk; spherical or diverging waves lead to distortions typical of near-field (Fresnel) diffraction. Observation must occur precisely at the focal plane of a converging lens placed after the circular aperture to capture the far-field pattern without additional phase variations. For verification, a pinhole serves as the circular aperture to simulate point-source diffraction. In a representative laboratory setup, a helium-neon laser beam passes through a circular aperture and is focused by a lens, yielding an Airy disk observable on a screen. Detection typically involves high-contrast imaging techniques, such as projection onto a white screen or photographic capture, to resolve the faint outer rings against the bright center. For naked-eye visibility without magnification, the angular size of the pattern must exceed about 1 arcminute, matching the approximate resolution limit of human vision under ideal conditions.

Influencing Factors

In optical systems, various imperfections and environmental factors can degrade the ideal Airy disk pattern, leading to blurring, asymmetry, or reduced visibility of its characteristic central disk and concentric rings. Aberrations, such as spherical and chromatic types, introduce wavefront errors that distort the diffraction pattern. Spherical aberration causes light rays from different zones of the aperture to focus at varying points along the optical axis, resulting in a blurred and expanded central disk with energy redistributed to the surrounding rings, thereby softening the overall image. Chromatic aberration exacerbates this by dispersing wavelengths differently, producing colored fringes around the Airy disk and further reducing contrast due to wavelength-dependent focal shifts. Defocus, another common aberration, shifts the pattern away from the focal plane, spreading the energy of the central disk and altering the positions of the rings without breaking radial symmetry, which can make the pattern appear larger and less defined. Atmospheric turbulence represents a significant external influence, particularly in astronomical observations, where temperature variations in the air create random refractive index fluctuations that warp incoming wavefronts. This turbulence produces a "seeing disk" that convolves with the Airy disk, effectively enlarging the point spread function and obscuring fine details; for instance, typical seeing conditions yield a disk size of about 1 arcsecond, far exceeding the diffraction-limited Airy disk radius of approximately 0.14 arcsecond for a 1-meter telescope aperture at visible wavelengths (λ ≈ 550 nm). Such distortions arise from eddies and thermal gradients in the atmosphere, which introduce phase delays that smear the pattern over time. Additional system imperfections, including dust particles or scratches on the aperture, can induce asymmetries in the Airy pattern by locally altering the aperture's uniformity, leading to uneven diffraction and irregular ring structures rather than the symmetric ideal. Partial coherence in the incident light, often from extended sources, reduces the contrast between the central disk and rings by broadening the effective point spread function, as the mutual coherence function deviates from the coherent case and diminishes high-frequency details in the image. Similarly, temperature gradients in the surrounding air or within optical components generate refractive index variations that warp wavefronts, causing further distortion and asymmetry in the observed pattern. To mitigate these degrading factors, adaptive optics systems employ deformable mirrors and wavefront sensors to dynamically correct aberrations and turbulence-induced distortions in real-time, restoring the Airy disk closer to its diffraction-limited form in applications like telescopes. In laboratory settings, controlled environments such as cleanrooms minimize dust and particulate contamination on optical surfaces, ensuring higher pattern fidelity during precise diffraction experiments.

Approximations and Modifications

Gaussian Profile Approximation

The Airy disk intensity profile, arising from diffraction through a circular aperture, can be effectively approximated by a Gaussian function in the central region for simplified analytical treatments in optical systems design and simulation. This approximation facilitates computations involving beam propagation, convolution operations, and energy distribution analyses, where the exact Airy pattern's oscillatory rings are negligible or averaged out. The method is particularly valuable in fields like laser optics and microscopy, enabling closed-form solutions that would otherwise require numerical integration of the Bessel-based Airy formula. The Gaussian approximation takes the form I(r) \approx I_0 \exp\left( -\frac{2 r^2}{w^2} \right), where I_0 is the peak intensity, r is the radial distance, and w \approx 0.42 \, \lambda / \mathrm{NA} is the $1/e^2 radius parameter, with \lambda denoting the wavelength and \mathrm{NA} the numerical aperture; this value of w is selected to align the Gaussian width with the Airy disk's first minimum radius of approximately $0.61 \, \lambda / \mathrm{NA}. The derivation typically employs a least-squares fit to the Airy intensity I(r) = I_0 \left[ 2 J_1(kr \mathrm{NA}/f) / (kr \mathrm{NA}/f) \right]^2 (with J_1 the first-order Bessel function of the first kind, k = 2\pi / \lambda, and f the focal length), constrained by peak normalization to ensure the central lobe matches closely; such fits yield relative errors below 5% within the core region up to about half the Airy radius. Alternatively, the approximation stems from the Fourier transform properties of optical systems: the Airy pattern is the Hankel transform of a uniform circular pupil function, but replacing the pupil with a Gaussian apodization (which has a Gaussian transform) provides a similar central profile while smoothing the sidelobes, aiding propagation modeling in paraxial optics. A notable quantitative match between the profiles is their energy containment: the Gaussian encloses approximately 86% of its total integrated power within the $1/e^2 radius w, compared to roughly 84% of the energy in the Airy disk's central lobe bounded by its first dark ring. This close correspondence supports the approximation's validity for applications assessing beam quality, such as computing the propagation factor M^2, where the diffraction-limited Airy pattern is modeled as an ideal Gaussian beam with M^2 = 1 to evaluate focusability and far-field divergence. The primary advantage lies in the Gaussian's mathematical tractability, allowing explicit evaluation of integrals for tasks like aberration tolerance analysis or mode coupling in laser systems, without the computational overhead of the Airy pattern's infinite series expansion.

Obscured and Apodized Patterns

In optical systems featuring a central obscuration, such as those in reflecting telescopes with a secondary mirror, the diffraction pattern deviates from the standard Airy disk produced by a clear circular aperture. The obscuration creates an annular aperture with obscuration ratio ε (the ratio of the obscured diameter to the full aperture diameter, typically 0 ≤ ε < 1), leading to a modified intensity distribution where energy is redistributed from the central maximum to the surrounding rings. This results in a contraction of the central disk's angular radius while the rings become relatively brighter, enhancing contrast in certain low-contrast scenarios but reducing peak central intensity. The intensity for an annular aperture is given by I(\theta) \propto \left[ \frac{J_1(x) - \epsilon^2 J_1(\epsilon x)/\epsilon}{(1 - \epsilon^2)} \right]^2, where x = \frac{\pi D \sin \theta}{\lambda}, D is the aperture diameter, \theta is the angular displacement from the optic axis, \lambda is the wavelength, and J_1 is the first-order Bessel function of the first kind. The first minimum (defining the central disk radius) occurs at a smaller x than the unobscured case (where it is at x \approx 3.83), shrinking the disk size; for example, at ε = 0.3, the first zero radius decreases by approximately 10-15% compared to ε = 0, while secondary ring intensities increase notably. This formula arises from the superposition of diffraction contributions from the outer annulus, normalized by the effective pupil area $1 - \epsilon^2. In Schmidt-Cassegrain telescopes, where ε ≈ 0.3 is common, this pattern improves resolution slightly for high-contrast objects but transfers about 20-30% more energy to the rings, potentially complicating faint companion detection. The Hubble Space Telescope exemplifies these effects, with its secondary mirror producing an obscuration ratio of approximately 0.12-0.15, which contracts the central Airy disk and brightens diffraction rings, influencing point-spread function (PSF) performance in high-resolution imaging. This modification aids in suppressing some stellar leakage in coronagraphic modes but requires careful calibration to mitigate ring-induced artifacts in exoplanet searches. Apodization involves applying an amplitude-modulating mask or weighting function to the pupil, such as Gaussian or super-Gaussian profiles, to suppress sidelobes in the Airy pattern at the expense of broadening the central lobe. For a Gaussian apodization A(r) = \exp(- (r / \sigma)^2 ), where r is the radial coordinate in the pupil and σ controls the taper, the resulting PSF exhibits exponentially decaying tails instead of oscillatory rings, reducing sidelobe levels by factors of 10-100 depending on the apodization strength. This technique trades off a modest increase in central disk size—typically 10-20% wider than the unobscured Airy disk—for enhanced dynamic range, as the broadened main lobe contains a higher fraction of total energy (e.g., >90% within the core for strong apodization). In coronagraphy for exoplanet imaging, apodization is particularly valuable, as it minimizes starlight leakage beyond the focal plane mask by eliminating Airy rings, enabling contrasts up to 10^{-10} near the star. Super-Gaussian profiles A(r) = \exp(- (r / \sigma)^n ) with n > 2 offer sharper transitions, further suppressing distant sidelobes while preserving inner working angle performance, though the disk broadening limits the minimum resolvable separation to about 2-3 times the classical Airy radius. These methods are implemented in space-based instruments like those on the James Webb Space Telescope precursors, balancing sidelobe suppression with minimal resolution loss.

Non-Circular Apertures

When the aperture deviates from circular symmetry, the resulting diffraction pattern becomes anisotropic, lacking the radial symmetry of the Airy disk. For an elliptical aperture with widths D_x along the minor axis and D_y along the major axis (D_y > D_x), the far-field intensity distribution approximates the product of two separable one-dimensional diffraction patterns, expressed as I(\theta_x, \theta_y) \propto \left[\mathrm{sinc}\left(\frac{\pi D_x \theta_x}{\lambda}\right)\right]^2 \left[\mathrm{sinc}\left(\frac{\pi D_y \theta_y}{\lambda}\right)\right]^2, where \theta_x and \theta_y are angular coordinates, \lambda is the wavelength, and the first minima occur at \theta_x = \lambda / D_x and \theta_y = \lambda / D_y. This yields a stretched central lobe along the major axis direction, with the pattern resembling a scaled and sheared version of the circular Airy pattern for low eccentricity. The resolution along each axis mirrors the circular case but scaled by the local dimension, with angular radii approximated as \theta_x = 1.22 \lambda / D_x and \theta_y = 1.22 \lambda / D_y, though the factor 1.22 derives from circular geometry and serves as a conventional analogy for defining the effective Airy disk extent in anisotropic cases. For near-circular ellipses, an effective diameter D_\mathrm{eff} = \sqrt{D_x D_y} provides a reasonable approximation for overall resolution scaling, treating the pattern as quasi-circular with this geometric mean dimension. For arbitrary non-circular shapes, the Fraunhofer diffraction pattern is generally the squared modulus of the Fourier transform of the pupil function, producing non-ring-like structures tailored to the aperture geometry. A square aperture of side length D, for instance, generates I(\theta_x, \theta_y) \propto \left[\mathrm{sinc}\left(\frac{\pi D \theta_x}{\lambda}\right)\right]^2 \left[\mathrm{sinc}\left(\frac{\pi D \theta_y}{\lambda}\right)\right]^2, forming a square central bright region bounded by minima at \theta_x = \theta_y = \lambda / D, surrounded by orthogonal lobes rather than concentric rings. In practical systems, such patterns influence performance in specialized optics. Adaptive optics employing deformable mirrors can introduce irregular effective pupils through wavefront shaping or partial obscurations, leading to axis-dependent resolution variations that must be modeled via the pupil's Fourier transform for optimal correction. Similarly, slit spectrographs utilize narrow rectangular apertures, where the sinc-like diffraction along the slit width limits spectral resolution, with the central spot's extent determining the minimum resolvable wavelength separation \Delta\lambda \approx \lambda / (D_\mathrm{slit} \cdot N), N being the number of grating grooves illuminated.

Comparisons

To Gaussian Beam Focusing

The diffraction-limited focus produced by a uniform illumination of a circular aperture results in an Airy pattern, characterized by a central bright disk surrounded by concentric sidelobes that contain approximately 16% of the total energy, with 84% confined within the central disk up to the first intensity minimum. In contrast, the fundamental transverse electromagnetic mode (TEM00) of a Gaussian beam exhibits a smooth intensity profile without sidelobes, with roughly 86% of its energy enclosed within the 1/e2 radius. This absence of sidelobes in the Gaussian beam provides a cleaner focal spot, reducing unwanted light scattering in applications such as laser processing or high-resolution imaging. During free-space propagation, Gaussian beams maintain their transverse profile shape indefinitely, characterized by a beam quality factor M2 = 1 for the ideal TEM00 mode, leading to predictable divergence governed by the beam parameter product. The focused waist radius w0 for such a beam, when the input 1/e2 diameter matches the aperture diameter D, is given by w0 = \frac{2 \lambda f}{\pi D}, where \lambda is the wavelength and f is the focal length. In comparison, the Airy pattern from a hard-edged uniform aperture diverges more rapidly, with the surrounding rings expanding and interfering, resulting in a non-self-similar propagation that degrades beam quality beyond the focal plane; its angular radius to the first minimum is 1.22 \lambda / D. In practical optical systems, single-mode optical fibers emit beams that closely approximate the Gaussian TEM00 mode due to the guided wave nature of the fundamental mode, enabling low-divergence outputs suitable for coupling into free-space optics. Conversely, multimode fibers or systems with hard-edged apertures, such as pinholes or uniform laser illuminators, produce patterns resembling the Airy disk, with pronounced sidelobes that can limit performance in precision applications. The beam quality factor M2 serves as a key metric in laser systems, where a value of 1 indicates ideal Gaussian behavior, while uniform aperture illumination yields an effective M2 greater than 1 due to diffraction losses into sidelobes. A primary trade-off in aperture illumination arises between central intensity and diffraction control: uniform illumination across the aperture maximizes the peak intensity at the focus by efficiently utilizing the full area, but it excites significant energy into sidelobes, broadening the effective spot and increasing sensitivity to aberrations. Gaussian apodization, by contrast, sacrifices some central intensity for sidelobe suppression and better propagation stability, making it preferable in systems requiring minimal divergence, such as fiber lasers or microscopy.

To Other Diffraction Phenomena

The Airy disk arises from the Fraunhofer diffraction through a circular aperture, producing a rotationally symmetric intensity pattern characterized by a central bright spot surrounded by concentric rings. In contrast, diffraction from a single rectangular slit yields a one-dimensional sinc² pattern, with intensity given by I(\theta) = I_0 \left( \frac{\sin \beta}{\beta} \right)^2, where \beta = \frac{\pi a \sin \theta}{\lambda} and a is the slit width; this results in linear fringes extending primarily along the direction perpendicular to the slit. The first minimum in the single-slit pattern occurs at an angular separation \theta = \lambda / a, defining a resolvable width of approximately \lambda / a, whereas the Airy disk's first minimum is at \theta = 1.22 \lambda / D (with D the aperture diameter), making the circular pattern broader for equivalent dimensions and concentrating more energy (about 84%) in the central disk compared to the slit's more distributed side lobes. Diffraction gratings, consisting of multiple parallel slits, generate sharp interference maxima at discrete orders governed by d \sin \theta = m \lambda (where d is the grating spacing and m the order), but these are modulated by the underlying single-slit envelope for rectangular grooves, leading to a sinc² modulation that suppresses higher orders. In optical systems with circular pupils, such as astronomical spectrographs employing echelle gratings (coarse, high-order blazed gratings for broad wavelength coverage), the overall diffraction envelope takes the form of the Airy pattern, which shapes the intensity distribution across spectral lines and limits resolution in the dispersion direction. This Airy modulation ensures that off-axis orders fade more gradually due to the pattern's radial symmetry, unlike the sharper cutoff in linear slit envelopes. A key specific case involves double-slit interference patterns confined within the Airy disk, as occurs when source separation is sub-resolution (less than the Rayleigh criterion of 1.22 λ / D); here, the interference fringes are enveloped by the disk's central lobe, enabling techniques like structured illumination for surpassing the classical diffraction limit. In Fourier optics, the Airy pattern represents the squared magnitude of the two-dimensional Fourier transform of the uniform circular pupil function, underscoring its role as the point spread function for ideal circular apertures in imaging systems. All such diffraction phenomena, including the Airy disk, embody the uncertainty principle in wave optics: confining light spatially through an aperture of size \Delta x inevitably broadens its angular spread \Delta \theta \approx \lambda / \Delta x, with the Airy pattern exemplifying the isotropic limit for circular geometries where momentum uncertainty manifests symmetrically in all directions.

References

  1. [1]
    Airy Disk explanation from Field Guide to Geometrical Optics - SPIE
    An aberration-free optical system does not image a point to a point. An Airy disk is produced having a bright central core surrounded by diffraction rings.Missing: definition authoritative
  2. [2]
    The Airy Disk and Diffraction Limit
    **Summary of Airy Disk Content:**
  3. [3]
    On the Diffraction of an Object-glass with Circular Aperture - ADS
    On the Diffraction of an Object-glass with Circular Aperture. Airy, G. B.. Abstract. Publication: Transactions of the Cambridge Philosophical Society.
  4. [4]
    Fundamental Aspects of Airy Disk Patterns - ZEISS
    May 22, 2024 · Learn about the basics of Airy disk patterns in microscopy, including their generation, size changes with illumination, and resolving power.Missing: authoritative | Show results with:authoritative
  5. [5]
    optics_airy_disk.html - UNLV Physics
    An Airy diffraction pattern is the diffraction pattern of a plane waves impinging on a circular aperture in a screen that is parallel to the wave fronts.
  6. [6]
    [PDF] Chapter 3 Diffraction
    The PSF, being the absolute value squared of the Fourier transform of the aperture, will thus not be precisely an Airy pattern, but a distorted Airy pattern.
  7. [7]
    Circular Aperture Diffraction - HyperPhysics
    When light from a point source passes through a small circular aperture, it does not produce a bright dot as an image, but rather a diffuse circular disc.
  8. [8]
    [PDF] Chapter 8: Diffraction [version 1008.1.K] - Caltech PMA
    In the nearby, Fresnel region, the aperture creates a beam whose edges will have the same shape and size as the aperture itself, and will be reasonably sharp ( ...
  9. [9]
    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 ...
  10. [10]
    July 1816: Fresnel's Evidence for the Wave Theory of Light
    Later he used those same equations to predict the interference patterns produced by two mirrors reflecting light. That became the basis for his 1818 treatise, ...
  11. [11]
    XXXI. Investigations in optics, with special reference to the ...
    Investigations in optics, with special reference to the spectroscope. Lord Rayleigh FRS. Pages 261-274 | Published online: 13 May 2009.
  12. [12]
    XV. On the theory of optical images, with special reference to the ...
    On the theory of optical images, with special reference to the microscope. Lord Rayleigh Sec. RS. Pages 167-195 | Published online: 08 May 2009.
  13. [13]
    [PDF] Quantum Optics Experiments in Space
    Indeed, the quantum theory, born at the beginning of the XX century to ... Airy disk given by a circular aperture [106]. The FFDP of a TIR corner cube has ...<|control11|><|separator|>
  14. [14]
    None
    Below is a merged summary of the Fraunhofer diffraction pattern derivation for a circular aperture, consolidating all relevant information from the provided segments. To maximize detail and clarity, I will use a structured format with tables where appropriate, followed by a narrative summary. The response retains all key details, including pupil functions, Fourier transforms, amplitude, intensity, energy in the central disk, variables, and useful URLs, while addressing discrepancies or variations across the segments.
  15. [15]
    [PDF] diffraction from a circular aperture
    the intensity distribution of Equation 13.15 is usually called the Airy pattern, and the central maximum within the first dark ring is called the Airy disc.
  16. [16]
    Two-Dimensional Fourier Optics
    In this limit, the interference/diffraction pattern appearing on the projection screen is entirely due to the phase differences between the spherical waves ...
  17. [17]
    Resolving power
    The radius r of the Airy Disc at the focal point of a lens is given by r = 1.22 λf/D. If all ray aberrations in an optical system can be eliminated, such that ...
  18. [18]
    Lecture 37: Diffraction on circular apertures
    This pattern was investigated by George Airy and the central bright spot is known as Airy disk ... 83.8% of the power is contained within the first dark ...
  19. [19]
    [PDF] space - NASA Technical Reports Server (NTRS)
    Unlike ground-based telescopes, the 2.4-m Space Telescope will possess and can effectively utilize an optical quality of such precision that its resolving power ...
  20. [20]
    Optimizing system resolution: a practical guide to matching lens and ...
    A common initial approach to pairing a lens and sensor is to match the radius of the lens's Airy disk to the sensor's pixel size. ... sensor is undersampling the ...
  21. [21]
    Diffraction Limited Photography: Pixel Size, Aperture and Airy Disks
    Diffraction Pattern. For an ideal circular aperture, the 2-D diffraction pattern is called an "airy disk," after its discoverer George Airy. The width of the ...
  22. [22]
    7.3. apodizing mask for telescopes
    Apodizing, or change of aperture's transmission properties, also affects intensity distribution within the diffraction pattern it generates.
  23. [23]
    Design, manufacturing, and testing of phase-induced amplitude ...
    While pupil apodization by creating a soft-edged pupil reduces the Airy rings, PIAA apodization modifies the pupil structure. For instance, our M1/M2 module ...
  24. [24]
    [PDF] Section 16 The Eye
    The iris is the stop of the eye. The pupil is the EP of the eye and has a typical diameter of about 4 mm, with a range of 2-8 mm. 1. 17 f. ϕ. ≡ ≈ mm. 23. R. V.
  25. [25]
    Facts and Figures concerning the human retina by Helga Kolb
    Range of pupil diameters: 1-8 mm. Refractive indices. Air: 1.000; Glass: 1.520; Water: 1.333; Cornea: 1.376. Optical power (diopters). Cornea: 43; Lens (relaxed): ...
  26. [26]
    [PDF] Optics and Image Quality in the Human Eye - Roorda Lab
    The PSF for a perfect optical system is the Airy disc, which is the Fraunhofer diffraction pattern for a circular pupil. ... Diffraction-limited Eye. 2.3 ...Missing: disk acuity
  27. [27]
    Visual Acuity by Michael Kalloniatis and Charles Luu - Webvision
    Jun 5, 2007 · Visual acuity is limited by diffraction, aberrations and photoreceptor density in the eye (Smith and Atchison, 1997).
  28. [28]
    L2S4 - JILA
    The HST has an angular resolution q = 0.05 arcseconds, 1200 times sharper. There are four main factors that might limit the sharpness of a telescope image. The ...
  29. [29]
    The Clinical Use of Vernier Acuity: Resolution of the Visual Cortex Is ...
    Oct 5, 2021 · Vernier acuity measures the ability to detect a misalignment or positional offset between visual stimuli, for example between two vertical lines ...
  30. [30]
    [PDF] Examining the Effects of Chromatic Aberration, Object Distance, and ...
    acuity is limited only by diffraction: a point source forms an Airy disk whose width is determined by the ratio of the diameter of the eye's pupil to the wave-.
  31. [31]
    [PDF] Astronomical Instruments
    Resolution: The maximum resolution of the human eye is typically one arc minute (1'). ... When stars ae closer Airy disks starts to overlap and making difficult ...
  32. [32]
    The Diffraction Limited Spot Size with Perfect Focusing
    Mar 13, 2017 · An important topic in diffraction-limited optics is the Airy Disk, which is defined as "the best focused spot of light that a perfect lens ...
  33. [33]
    Experimental study of spatial coherence diffraction based on full ...
    Similar to the well-known Airy disk stemming from optical diffraction, we studied the spatial coherence function near the coherence focal plane on the analogy ...
  34. [34]
    Why Should We Care About Spot Size in Fiber Laser Cutting
    Spot size refers to the diameter of the laser beam at its focus point. In other words, it is the area where the laser's energy is concentrated onto the material ...
  35. [35]
    [PDF] Diffraction Patterns Produced by Focused Laser Beams - RAND
    When the output beam is significantly truncated or obscured, the diffraction pattern is no longer gaussian; rather, it resembles an Airy pattern, i.e., a bright ...<|separator|>
  36. [36]
    Quick guide on laser beam divergence measurement - Gentec-EO
    Jul 11, 2023 · A Gaussian laser beam is said to be diffraction-limited when the measured divergence is close to θ0. To achieve the best aiming performance, ...
  37. [37]
    Single atom in a superoscillatory optical trap - Nature
    Jun 29, 2023 · The images are fitted with an Airy disk pattern, to extract the Abbe radius. For a diffraction-limited spot, the Abbe radius is μ , with ...Missing: LIDAR | Show results with:LIDAR
  38. [38]
    Super-resolution single-photon imaging at 8.2 kilometers
    Jan 29, 2020 · ... diffraction limit (Airy disk diameter). For single-photon LiDARs, an efficient way is to match the pixel size of the receiver with the point ...
  39. [39]
    Demonstration of the Airy disk using photography and simple light ...
    Nov 1, 2015 · A procedure for observing diffraction from a circular aperture using photography equipment and commonly available light sources is presented.
  40. [40]
    Fraunhoffer Diffraction - LNMIIT UG Physics Lab
    We can observe Fraunhoffer diffraction experimentally by using a collimated light source and (i) placing the viewing screen at the focal plane of a convex lens ...
  41. [41]
    Airy's disk: Pinhole diffraction using laser beam.
    Airy's disk is a circular diffraction pattern created by passing monochromatic laser light through a pinhole, then enlarged with a diverging lens.Missing: lab plane focal
  42. [42]
    The human eye - Amateur Telescope Optics
    For 0.55μ wavelength, the true angular size of the Airy disc formed by the objective is 4.6/D in arc minutes, for aperture diameter D in mm. The same relation ...
  43. [43]
    6.4. Diffraction pattern and aberrations - Amateur Telescope Optics
    The main and most damaging effect is brightening of the rings, expanding and softening the point-object images. The enlargement of the central disc itself is ...
  44. [44]
    Get to know the Optical Aberrations present in your Image
    Below are some examples of aberrations, viewed individually, that can be present in final imagery. Each example depicts a diffraction-limited airy disk (the ...
  45. [45]
    Effect of atmospheric turbulence on the telescope image (seeing error)
    The central maxima of PSF expanded by turbulence (atmospheric PSF) is called the seeing disc, and its FWHM the seeing. ... Airy disc. Verbal description of the ...Missing: disk
  46. [46]
    Atmospheric Seeing - RASC Calgary Centre
    Nov 5, 2018 · The Airy disk gets smaller as the telescope aperture increases. making it difficult to see either clearly.
  47. [47]
    The effects of optical coatings and glass inhomogeneity on image ...
    Refractive index inhomogeneity (RII) is caused by variations in glass density resulting from the production process. The pattern can have any type of form, ...
  48. [48]
  49. [49]
    Ambiguity of the Transfer Function with Partially Coherent Illumination
    It is shown that the significant parameter is the ratio of coherence interval to the diameter of the Airy disk (or diffraction spot) of the imaging lens. It is ...
  50. [50]
    [PDF] study of passive optical techniques for detecting clear air turbulence
    Remote detection of temperature gradients ahead of an aircraft by infrared radiometry should also be actively persued. Although it has been shown in this report ...
  51. [51]
    [PDF] Flow, Heat Transfer, and Wavefront Distortion - OSTI
    Jan 16, 1989 · As the surface heat flux increases, either the temperature gradients get so steep that thermo-mechanical stresses eventually break the disk or ...
  52. [52]
    Correction of static optical errors in a segmented adaptive optical ...
    Adaptive optical systems are designed to compensate for wave-front errors caused by atmospheric turbulence. In addition, they may also correct for ...Missing: mitigate | Show results with:mitigate
  53. [53]
    Principle of Cleanroom - PMC - PubMed Central - NIH
    Cleanroom is the most important and most representative measure to create the clean microenvironment with air cleaning technology.Missing: Airy | Show results with:Airy
  54. [54]
    Airy disk diffraction pattern: comparison of some values of f/No. and ...
    The Airy disk pattern's radii of zeros and maxima are calculated for different f/No. and obscuration ratios, with formulas provided. Tables show these radii ...
  55. [55]
    Effects of the telescope central obstruction
    Central obstruction does lower contrast for both, sinusoidal and square-wave patterns, but it degrades the low-contrast resolution limit only with the former.Missing: Hubble | Show results with:Hubble
  56. [56]
    [PDF] Searching for Faint Companions to Nearby Stars with the Hubble ...
    The first two terms in Eq. (1) represent the annular pupil created by the PC secondary-mirror obscuration. The third term describes the obscuration.
  57. [57]
    Apodization pupils: design and performance (review) - IOPscience
    The magnitude of reduction of the sidelobes was found to increased by the amplitude apodization across the slit function. For apodized circular pupil functions, ...Missing: disk | Show results with:disk
  58. [58]
    An Introduction to Diffraction. Part II: The Far Field - Sage Journals
    The diffracted ampli- tude would be described by a sinc ... Conversion from a circular to an elliptical aperture will produce an el- liptically shaped diffraction ...
  59. [59]
    [PDF] Spectrographs and Spectroscopy
    • The single slit diffraction pattern modifies this by affecting the heights ... For slit spectral, depends on slit width and grating choice ...
  60. [60]
    [PDF] 2.4 Paraxial Wave Equation and Gaussian Beams
    Thus, there is a certain fraction of power within a certain radius of the beam. P(r < w(z)). P. = 0.86,. (2.236). P(r < 1.5w(z)). P. = 0.99. (2.237). Beam ...
  61. [61]
    Gaussian beams - RP Photonics
    Definition of Gaussian Beams​​ Here, the beam radius is the distance from the beam axis where the intensity drops to 1 / e 2 (≈ 13.5%) of the maximum value. If ...
  62. [62]
    Gaussian Beam Optics - Newport
    The “focal spot size” given by our simple equation would be 200 times larger, or 1.6 mm, 60% larger than the original beam! Obviously, something is wrong. The ...
  63. [63]
    [PDF] Gaussian Beam Optics
    Far-field divergence (the number quoted in laser specifications) must be measured at a distance much greater than zR. (usually >10#zR will suffice). This is a ...Missing: LIDAR | Show results with:LIDAR
  64. [64]
  65. [65]
    [PDF] Diffraction occurs when light waves is passed by an aperture/edge
    Diffraction occurs when light waves pass by an aperture/edge, and each point on the wavefront acts as a source of another circular wave.
  66. [66]
    Why do higher orders in diffracted light show up as shades of one ...
    Sep 17, 2024 · Other wavelengths satisfy the grating equation as you go farther off axis, but their intensity decreases due to the Airy pattern, so you see an ...
  67. [67]
    Characterization of Diffraction Grating Imaging Quality - Newport
    In Concave Diffraction Gratings, we formulated the optical imaging properties of a grating system in terms of wavefront aberrations.
  68. [68]
    (PDF) The uncertainty principle in classical optics - ResearchGate
    Aug 9, 2025 · In the focal plane, the distance between the focused spots is f , which must be greater than the Airy disk radius of 1.2 f /D if the individual ...
  69. [69]
    [PDF] Chapter 7 Diffraction - Caltech PMA
    The finite size causes different pieces of the source to have their Fresnel rings centered at slightly different points in the aperture plane, and this causes ...