Fact-checked by Grok 2 weeks ago

van Cittert–Zernike theorem

The van Cittert–Zernike theorem is a foundational result in optical that establishes the mutual between two points in a wave field illuminated by a quasi-monochromatic, spatially incoherent extended source as the normalized of the source's intensity distribution in the far-field (Fraunhofer) approximation. This theorem quantifies how partial spatial emerges upon propagation from an initially incoherent source, with the degree of decreasing with increasing separation between the observation points and source size. Named after Dutch physicists Pieter Hendrik van Cittert and Frits Zernike, the theorem originated from van Cittert's 1934 analysis of wave field spectral properties influenced by source geometry, which laid the groundwork for linking source distribution to propagation. Zernike provided a more accessible proof in while introducing the concept of the complex degree of as the normalized mutual intensity, enabling practical applications in . These contributions revitalized interest in partial during the early , bridging classical wave with statistical descriptions of light. The theorem assumes a planar, quasi-monochromatic source with incoherent emission (zero mutual across the source plane), free-space propagation without or , and observation in the far field where the Fresnel number is much less than unity. Mathematically, for a source I(\mathbf{\xi}) in the source plane, the complex degree of \gamma_{12} at points \mathbf{r}_1 and \mathbf{r}_2 in the observation plane separated by distance z is given by \gamma_{12} = \frac{\int I(\mathbf{\xi}) \exp\left[-i \frac{2\pi}{\lambda z} (\mathbf{r}_2 - \mathbf{r}_1) \cdot \mathbf{\xi}\right] d\mathbf{\xi}}{\int I(\mathbf{\xi}) d\mathbf{\xi}}, revealing the relationship. Extensions to vector electromagnetic fields maintain this form for the spectral degree of while preserving source-plane . Applications span physical optics, including the design of partially coherent illumination systems for microscopy and lithography, where controlled coherence reduces speckle and improves resolution. In radio astronomy, the theorem underpins aperture synthesis interferometry, allowing visibility measurements between antennas to reconstruct extended celestial sources via inverse Fourier transform, as in the Very Large Array. Quantum and multiphoton generalizations further apply to entangled photon propagation and matter-wave interferometry.

Introduction

Historical Background

The van Cittert–Zernike theorem emerged from early 20th-century investigations into the propagation of light from incoherent sources, with foundational contributions from Dutch physicists Pieter Hendrik van Cittert and Frits Zernike. Van Cittert's work in 1934 addressed the statistical distribution of light vibrations in a plane illuminated directly or via a by an extended source, providing an initial formulation of the relationship between source intensity and field . This analysis was motivated by challenges in understanding patterns and effects observed in astronomical contexts, such as stellar , where resolving the angular structure of distant stars required modeling the partial of light from extended celestial sources. In 1938, Frits Zernike provided a simpler and more accessible proof of van Cittert's result, simplifying and generalizing the approach by introducing the concept of the degree of as a quantitative measure of spatial in wave fields. Zernike's formulation appeared in his seminal paper on the application of ideas to optical instruments, which built directly on van Cittert's statistical framework while emphasizing practical implications for imaging and interferometric systems. Although Zernike's broader research in , including his development of (recognized with the in 1953), was distinct, his 1938 contributions to theory complemented these efforts by providing tools to analyze partially coherent illumination in microscopic and telescopic observations. The theorem, combining these insights, gained prominence in the mid-20th century as a fundamental principle in the emerging field of partial coherence theory, influencing subsequent advancements in and . Key publications, including van Cittert's Die wahrscheinliche Schwingungsverteilung in einer von einer Lichtquelle entweder direkt oder mittels einer Linse beleuchteten Ebene (Physica 1, 201–210, 1934) and Zernike's The Concept of Degree of Coherence and Its Application to Optical Instruments (Physica 5, 785–795, 1938), established the theorem's rigorous basis and spurred its integration into theoretical frameworks for wave propagation and instrument design.

Statement of the Theorem

The van Cittert–Zernike theorem states that, for an incoherent quasi-monochromatic extended source, the complex degree of coherence between two points in a far-field plane is equal to the normalized of the source's intensity distribution. This result, originally derived by Pieter Hendrik van Cittert in 1934 and reformulated more accessibly by Frits Zernike in 1938, establishes a direct link between the spatial structure of the source and the degree of spatial coherence in the propagated field. Mathematically, the mutual coherence function \Gamma(P_1, P_2) is given by \Gamma(P_1, P_2) = \frac{\int I(\xi) \exp\left[-i \frac{2\pi}{\lambda z} (\xi \cdot (P_1 - P_2))\right] d\xi}{\int I(\xi) d\xi}, where I(\xi) denotes the distribution across the source plane in coordinates \xi, \lambda is the , z is the propagation to the observation plane, and P_1, P_2 are position vectors of the two points in that plane. The normalization ensures that \Gamma(P, P) = 1 for any point P, reflecting full with itself. This formulation highlights how the theorem quantifies spatial as inversely related to the angular extent of the source: larger sources produce shorter coherence lengths in the observation plane, with the phase term in the integrand encoding the interferometric dependence on point separation.

Core Concepts

Mutual Coherence Function

The mutual coherence function, introduced by Emil Wolf in , serves as a fundamental quantity in the statistical description of partially coherent optical fields. It is defined as the ensemble average of the product of the complex amplitudes of the field at two points P_1 and P_2 separated by a time delay \tau: \Gamma(P_1, P_2, \tau) = \langle U(P_1, t) \, U^*(P_2, t + \tau) \rangle, where U(P, t) represents the complex of at position P and time t, the asterisk denotes complex conjugation, and \langle \cdot \rangle denotes the ensemble average over the . This function captures both spatial and temporal correlations in . For the analysis of spatial , which is central to many optical phenomena, the case \tau = 0 is typically considered, yielding the mutual at equal times \Gamma(P_1, P_2) = \langle U(P_1, t) \, U^*(P_2, t) \rangle. A related quantity is the complex degree of spatial , which the mutual coherence function to provide a dimensionless measure bounded between zero and one in . It is given by \gamma(P_1, P_2) = \Gamma(P_1, P_2) / \sqrt{I(P_1) I(P_2)}, where I(P) = \Gamma(P, P) is the (or mean ) at point P. This distinguishes the degree of from the mutual coherence function itself, as the latter includes amplitude information through the intensities, while \gamma isolates the . Physically, the mutual coherence function quantifies the strength and relationship between the vibrations of the field at the two points; the modulus |\gamma(P_1, P_2)| equals 1 for fully coherent fields (perfect locking) and approaches 0 for completely incoherent fields (no predictable relation). In the framework of the van Cittert–Zernike theorem, the mutual function plays a pivotal role by connecting the statistical properties of an incoherent source—characterized primarily by its —to the structure of the field in the far-field observation plane. This linkage enables the prediction of partial patterns emerging from extended sources without assuming full .

Intensity of the Source

The van Cittert–Zernike theorem relies on the assumption of an incoherent extended source, conceptualized as a collection of independent point emitters distributed across the source plane, where each emitter contributes light with random, uncorrelated phases relative to others. This model captures the statistical nature of thermal or sources, such as stars or lamps, where phase relationships between distinct points are absent due to the random processes. The I(\xi), with \xi denoting the transverse coordinate in the source plane, is defined as the ensemble average of the squared field amplitude, I(\xi) = \langle |U(\xi)|^2 \rangle, representing the local without fixed information. Typical spatial distributions of I(\xi) include uniform disks, which model circularly symmetric sources like pinholes or unresolved stellar disks, and Gaussian profiles, often used for approximations of extended atmospheric or thermal emitters. The angular size \theta of such a source, subtended at the observation plane, inversely scales the spatial coherence length to roughly \lambda / \theta, where \lambda is the wavelength, highlighting how larger sources reduce coherence over finite apertures. To ensure physical consistency, the profile is normalized such that the over the source area yields the total emitted power, \int I(\xi) \, d\xi = P, preserving in the model. In contrast to coherent sources, which feature deterministic relations across the entire and thus propagate with preserved full , the lack of fixed phases in incoherent sources leads to partial in the far field, as measured by the mutual coherence function.

Mathematical Derivation

Propagation Integral

The propagation of the optical field from an extended incoherent source to an observation plane is fundamentally described by the Huygens-Fresnel principle, which models the field as a superposition of secondary wavelets emanating from each point in the source plane. The complex amplitude U(\mathbf{P}) at an observation point \mathbf{P} located at a distance z along the is given by the U(\mathbf{P}) = \frac{1}{i \lambda z} \iint U(\boldsymbol{\xi}) \frac{\exp(i k r)}{r} \, d^2 \boldsymbol{\xi}, where \boldsymbol{\xi} denotes coordinates in the source plane, \lambda is the , k = 2\pi / \lambda is the wave number, and r = |\mathbf{P} - \boldsymbol{\xi}| is the from the source point to the observation point. This formulation arises from the Kirchhoff diffraction theory and assumes the source plane acts as an for wave propagation. In the far-field regime, where z \gg |\boldsymbol{\xi}| and z \gg |\mathbf{P}|, the denominator r is approximated as z, simplifying the factor while retaining the information in the exponential term. For incoherent sources, the mutual function \Gamma(\mathbf{P}_1, \mathbf{P}_2; \tau) = \langle U(\mathbf{P}_1, t) U^*(\mathbf{P}_2, t + \tau) \rangle at two observation points \mathbf{P}_1 and \mathbf{P}_2 is derived from the of the source fields, assuming statistical independence between different source points. Since the source intensity distribution I(\boldsymbol{\xi}) determines the incoherent nature, with the source mutual being \Gamma_s(\boldsymbol{\xi}_1, \boldsymbol{\xi}_2; \tau) = I(\boldsymbol{\xi}_1) \delta(\boldsymbol{\xi}_1 - \boldsymbol{\xi}_2) (neglecting temporal dependence for quasi-monochromatic light), the propagated becomes \Gamma(\mathbf{P}_1, \mathbf{P}_2) = \iint I(\boldsymbol{\xi}) h(\mathbf{P}_1, \boldsymbol{\xi}) h^*(\mathbf{P}_2, \boldsymbol{\xi}) \, d^2 \boldsymbol{\xi}, where h(\mathbf{P}, \boldsymbol{\xi}) is the complex propagation kernel from the Huygens-Fresnel integral, incorporating the amplitude and phase contributions from each source element. This expression captures how partial coherence emerges in the observation plane due to the superposition of waves from the extended source. Under the paraxial approximation, valid for small angles where the observation points lie near the optical axis, the distance r is expanded as r \approx z + \frac{|\mathbf{P} - \boldsymbol{\xi}|^2}{2z}, leading to a quadratic phase factor in the kernel: \exp(i k r) \approx \exp(i k z) \exp\left[i \frac{\pi}{\lambda z} |\mathbf{P} - \boldsymbol{\xi}|^2 \right]. This approximation neglects higher-order terms in the obliquity factor and spherical wave curvature, focusing on the dominant Fresnel diffraction effects while maintaining accuracy for beam-like propagation. In the Fraunhofer limit, applicable when the observation distance satisfies z \gg \frac{k |\boldsymbol{\xi}|^2}{2} (ensuring negligible quadratic phase variation across the source), the propagation simplifies to a decomposition into plane waves, where the kernel's phase becomes linear in the transverse coordinates, facilitating angular spectrum representations without altering the integral structure.

Fourier Transform Relation

Under far-field conditions, where the observation plane is sufficiently distant from the source such that the Fresnel approximation holds and the source subtends a small , the propagation for the mutual function Γ(P₁, P₂) simplifies significantly. The phase term in the , arising from the difference in path lengths, approximates to exp[-i (2π/λz) ξ · ΔP], where λ is the , z is the propagation distance, ξ is the source coordinate vector, and ΔP = P₁ - P₂ is the separation vector in the observation plane. This exponential form directly identifies the mutual as proportional to the two-dimensional of the source intensity distribution I(ξ) evaluated at the spatial frequency f = ΔP / (λz). The normalized complex degree of coherence γ(P₁, P₂) = Γ(P₁, P₂) / √[I(P₁) I(P₂)], where I(P) denotes the at point P, then takes the explicit form \gamma(\mathbf{P_1}, \mathbf{P_2}) \approx \frac{\iint I(\boldsymbol{\xi}) \exp\left[-i \frac{2\pi}{\lambda z} \boldsymbol{\xi} \cdot \Delta \mathbf{P}\right] d^2 \boldsymbol{\xi}}{\iint I(\boldsymbol{\xi}) d^2 \boldsymbol{\xi}}, with the denoted in the numerator. This relation establishes the core mathematical link of the theorem, showing that partial in the plane mirrors the Fourier spectrum of the incoherent source . This Fourier transform interpretation reveals key physical insights into spatial coherence. The transverse coherence length, defined as the separation ΔP over which |γ| drops to 1/e, scales inversely with the angular size of the source: larger sources produce shorter coherence lengths due to broader Fourier spectra. For a circular source of uniform intensity and radius a, the coherence pattern |γ(ΔP)| follows the Airy disk function J₁(π ΔP a / (λ z)) / (π ΔP a / (λ z)), analogous to the diffraction pattern from a circular aperture, with the first zero at ΔP ≈ 1.22 λ z / (2a). The spatial frequency argument f = ΔP / (λ z) further connects the source's angular extent θ ≈ a / z to the coherence scale, emphasizing how propagation distance z amplifies coherence for extended sources.80203-2)

Assumptions and Limitations

Incoherence and Quasi-Monochromatic Conditions

The van Cittert–Zernike theorem relies on the fundamental assumption that the source is spatially , whereby emissions from distinct points on the source occur independently with no fixed phase relationships between them. This statistical independence is modeled by δ-correlated phases across source elements, ensuring that the mutual function at the source plane simplifies to a delta function, Γ(ξ₁, ξ₂, τ=0) = I(ξ₁) δ(ξ₁ - ξ₂). Consequently, the total intensity in the observation plane arises solely from the incoherent superposition of contributions from each source point, given by I_{\text{total}}(\mathbf{r}) = \int I(\xi) \, d\xi, where I(\xi) is the intensity distribution over the source coordinate \xi, and no cross-interference terms appear due to the lack of phase correlations. This additivity is essential for deriving the Fourier transform relationship between the source intensity and the far-field coherence, as any residual coherence at the source would introduce unaccounted phase factors that invalidate the theorem's predictions. Historically, this incoherence originated in van Cittert's analysis of stellar sources, which he modeled as extended emitters where random processes produce uncorrelated from different surface elements, mimicking blackbody-like behavior without spatial locking. Zernike later formalized this in the context of partial , emphasizing its role in optical instrumentation for incoherent sources. For coherent sources, such as lasers, violating this disrupts the , as the pre-existing prevents the propagation-induced buildup central to the result.90026-4)80203-2) Additionally, the theorem presupposes quasi-monochromatic conditions, where the source satisfies Δλ ≪ λ, with the relative Δλ/λ typically much less than 1%. This narrow width enables a scalar wave approximation and permits time-averaging of the function over fluctuations slower than the optical period, treating the field as effectively monochromatic. The is thus represented as S(\omega) \approx S(\omega_0) \delta(\omega - \omega_0), concentrated around a central ω₀, which ensures that path-length differences in do not introduce significant within the time. Polychromatic sources with broader spectra lead to smearing of the pattern, as varying wavelengths produce misaligned components, thereby limiting the theorem's precision in relating source structure to observed .

Far-Field and Geometric Approximations

The van Cittert–Zernike theorem relies on the far-field condition, where the observation distance z greatly exceeds the square of the source size D divided by the \lambda, i.e., z \gg \frac{D^2}{\lambda}. This Fraunhofer ensures that incoming wavefronts from the source can be approximated as plane waves, simplifying the to a relationship between source intensity and mutual . A key geometric assumption is that the source subtends a small angular size \theta \ll 1 radian at the observation plane, allowing the ratio of source coordinate \xi to distance z to remain approximately constant across the source extent. This small-angle approximation linearizes the phase differences in the propagation, enabling the coherence function to directly relate to the source's angular intensity distribution without higher-order curvature terms. The theorem assumes a two-dimensional, planar extended source, neglecting any depth variations or three-dimensional structure that could introduce additional gradients. This simplification projects the source onto a flat distribution in the transverse plane, facilitating the two-dimensional formulation. Propagation occurs through a homogeneous medium that is isotropic and non-absorbing, free from , , or aberrations that could distort the wavefronts. This assumption maintains constant wave number k = 2\pi / \lambda throughout the path, preserving the direct mapping from source to . These approximations limit the theorem's validity: in the near-field (Fresnel) regime, where z is comparable to or less than D^2 / \lambda, quadratic phase terms must be retained, requiring more complex models. Similarly, sources with large extents violate the small-angle condition, leading to reduced visibility and the need for generalized formulations.

Extensions and Generalizations

Hopkins' Formula

In 1953, Harold Hopkins formulated a of the van Cittert–Zernike theorem to describe the propagation of through optical imaging systems that include finite apertures, aberrations, and other transmission properties. This extension builds on earlier concepts by incorporating the system's response to contributions from an extended incoherent source, enabling rigorous analysis of under realistic conditions. Hopkins' formula expresses the mutual coherence function \Gamma(P_1, P_2) between two points P_1 and P_2 in the observation plane as \Gamma(P_1, P_2) = \frac{1}{N} \iint I(\xi) \, t(\xi, P_1) \, t^*(\xi, P_2) \, d\xi, where I(\xi) is the intensity distribution across the source coordinates \xi, t(\xi, P) is the complex transmission function of the optical system from source point \xi to observation point P (encoding diffraction, aberrations, aperture effects, and propagation phases), t^* denotes its complex conjugate, and N is a normalization constant ensuring \Gamma(P, P) = I(P). The transmission functions t effectively define a coherent transfer function that modulates how source incoherence is transformed into spatial coherence in the image. This formulation accounts for the influence of imperfections on propagation, distinguishing it from the ideal free-space case of the original theorem. In the incoherent limit where t(\xi, P) = 1 for all points, ' expression recovers the standard van Cittert–Zernike relation via a of the source intensity. ' work on integrals laid foundational tools for , and the formula has proven essential for evaluating partial in instruments like and telescopes, where finite apertures limit the degree of across the field. For instance, it quantifies how source size and aberrations degrade image contrast in high-resolution .

Vector and Quantum Variants

The vector extension of the van Cittert–Zernike theorem addresses the of electromagnetic fields from spatially incoherent sources, generalizing the scalar case to account for effects. In 2009, Setälä, Tervo, and Friberg formulated the theorem using the 2×2 cross-spectral to characterize the correlation properties of the components E_x and E_y. This matrix describes the mutual between components in the far field as the of the source intensity distribution, with the degree of remaining invariant during free-space . For partially polarized sources, the theorem was further extended in 2013 to incorporate , enabling a description of both one-point and two-point properties. Tervo et al. established a relation between the source-plane and the far-zone , allowing the far-field properties of fields from incoherent sources—such as those with annular apertures—to be analyzed in terms of traditional metrics. This approach reveals how partial influences spatial patterns in the propagated field. In the quantum domain, a multiphoton variant of the van Cittert–Zernike theorem was introduced in 2023 by You et al., extending the classical framework to describe the propagation of entangled states without light-matter interactions. This quantum version uses a beam coherence-polarization matrix to model higher-order functions for multiphoton systems, such as biphotons generated via . For entangled photons, it predicts scattering-induced correlations that enhance second-order g^{(2)}, with values exceeding 1.5 at zero detector separation, applications in and sub-shot-noise sensing. The theorem highlights how free-space propagation modifies and bunching in scattering media. A related generalization concerns intensity correlations, where the theorem is adapted for the expectation value \langle I(\mathbf{P}_1) I(\mathbf{P}_2) \rangle between two points in the far field. In 2017, Shirai derived this form assuming circular Gaussian statistics in the source plane, linking the autocovariance to the source's noise power spectrum via a relation. This extension aligns with the principles of Hanbury Brown–Twiss interferometry, facilitating source size estimation from measurements, and also applies to quantized fields in coherent states exhibiting Poissonian fluctuations. These vector and quantum variants have gained recent relevance in wide-field radio , particularly for instruments like the (LOFAR) and the (SKA). In 2009, Carozzi and Karlsson developed a generalized van Cittert–Zernike theorem valid for partially polarized sources over arbitrary fields of view, correcting direction-dependent effects in visibility measurements and improving imaging accuracy for extended sources in these arrays.

Applications

Interferometry and Imaging

The van Cittert–Zernike theorem underpins techniques in by establishing that the complex measured between two antennas separated by baseline b is the of the incoherent source's intensity distribution I(ξ), where ξ represents angular coordinates on the sky. Specifically, the V(u) at u = b/λ (with λ the ) samples the Fourier plane, allowing reconstruction of I(ξ) through an inverse of the collected visibilities from multiple baselines. V(u,v) = \iint I(l,m) \, e^{-2\pi i (u l + v m)} \, dl \, dm This relationship enables synthesis of large effective apertures from arrays of smaller dishes, as demonstrated by the (VLA) in , which uses 27 antennas with maximum of 36 km to achieve resolutions down to 0.04 arcseconds at centimeter wavelengths. In stellar , the theorem sets the resolution limit via the spatial , where the minimum resolvable angular source size θ satisfies θ ≈ λ/(2b), with b the interferometer ; visibility fringes vanish when b exceeds this scale for extended sources. Michelson stellar exploits this to measure stellar diameters by observing the at which fringe contrast drops, providing direct angular size estimates for stars like . For modern wide-field arrays such as the () and (), extensions of the theorem accommodate extended, non-point-like sources across broad fields of view, enabling accurate imaging of diffuse structures like galactic emissions without the narrow-field approximations of classical formulations. In optical imaging applications, such as under partial , the theorem relates the illumination size to the degree of spatial in the object plane, where larger sources diminish mutual and reduce image by suppressing effects. The factor σ (source size normalized to the ) quantifies this, with σ = 1 yielding a contrast factor of 0.610 and narrowing the resolution limit by approximately 15% compared to incoherent limits. In , the theorem guides the use of partial in projection systems, such as steppers, to control σ and minimize unwanted effects like fringing, improving pattern fidelity in semiconductor manufacturing.

Coherence in Lasers and Modern Optics

In free-electron lasers (FELs), the van Cittert–Zernike theorem enables the of partial spatial from the finite size of the undulator source, which is essential for predicting beam quality and diffraction patterns in applications. For instance, at the Linac (LCLS), the theorem models how source angular distribution convolves with object scattering to reduce fringe contrast in , allowing techniques to recover high-resolution structures like protein nanocrystals with pulse durations of 10–300 fs and up to 10¹³ photons per pulse. This approach has advanced post-2010 FEL capabilities, such as serial femtosecond crystallography, by quantifying transverse lengths that must exceed twice the sample diameter for accuracy. The theorem also informs adaptive optics systems by modeling how atmospheric turbulence degrades spatial , guiding correction in ground-based telescopes to achieve near-diffraction-limited performance. In these setups, the van Cittert–Zernike relation describes the mutual function as the of the source intensity, distorted by turbulence-induced phase errors quantified by the r_0 (e.g., 0.2 m in visible wavelengths), enabling real-time compensation via deformable mirrors to minimize speckle and restore over apertures up to 8 m in diameter. Such modeling supports applications in large telescopes like the Keck Observatory, where it predicts isoplanatic angles for multi-conjugate , improving resolution for faint astronomical sources. In pulse-echo measurements, the van Cittert–Zernike theorem has been extended to and acoustics, predicting the spatial of backscattered fields from random media insonified by pulses, which is crucial for characterizing in . This generalization, developed in 1991, describes how the of the received depends on the scatterer distribution, analogous to optical propagation, and remains relevant to modern where light-induced from biological tissues exhibit similar partial behaviors for enhanced in deep-tissue visualization. Recent advances in speckle correlation techniques for through scattering media leverage the theorem to estimate correlation lengths, enabling recovery of hidden object sizes without direct . In these methods, the van Cittert–Zernike relation links intensity correlations in the far field to the source's angular spectrum, allowing depth-resolved by analyzing speckle decorrelation (e.g., via axial ) in turbid environments like biological tissues, with applications in non-invasive achieving sub-millimeter resolution. Quantum extensions of the van Cittert–Zernike theorem apply to multiphoton , where it governs the of nonclassical correlations in entangled photon pairs or higher-order states, improving signal-to-noise in biological of scattering samples. This multiphoton quantum version predicts how scattering media alter two-photon interference patterns, enabling super-resolution techniques in nonlinear for labeling-free visualization of cellular structures, such as neuronal networks, by exploiting reduced compared to classical sources. Further generalizations extend to matter-wave , where the theorem describes spatial of de Broglie waves from incoherent atomic or sources, facilitating high-resolution and interferometric measurements in atom experiments, such as with cold atom clouds or beams.

References

  1. [1]
  2. [2]
    On the van Cittert-Zernike theorem | Optical and Quantum Electronics
    The van Cittert-Zernike theorem states that the spatial coherence over a space illuminated by an incoherent extended source is described by the Fourier ...
  3. [3]
    generalized measurement equation and van Cittert-Zernike theorem ...
    Abstract. We derive a generalized van Cittert-Zernike (vC-Z) theorem for radio astronomy that is valid for partially polarized sources over an arbitrarily.
  4. [4]
    Multiphoton quantum van Cittert-Zernike theorem - Nature
    May 26, 2023 · Here, we introduce the quantum van Cittert-Zernike theorem to describe the scattering and interference effects of propagating multiphoton systems.
  5. [5]
  6. [6]
    The concept of degree of coherence and its application to optical ...
    The maximum visibility of the interferences obtainable from two points in a wave field is defined as their degree of coherence γ.
  7. [7]
    Zernike's research
    Sep 19, 2023 · Zernike's work helped awaken interest in coherence theory, the study of partially coherent light sources. In 1938 he published a simpler ...<|control11|><|separator|>
  8. [8]
    [PDF] Quantitative coherence and application to x-ray beam lines
    • Contrast transfer, reciprocity, influence of coherence on resolution. • Coherence and Zernike phase contrast, Wigner phase contrast. • Are microscopes flux ...
  9. [9]
    The van Cittert-Zernike theorem for electromagnetic fields
    One of the central results of optical coherence theory is the van Cittert-Zernike theorem which shows that even completely incoherent source can generate a ...
  10. [10]
    [PDF] The Van Cittert-Zernike Theorem - Princeton University
    The Van Cittert-Zernike Theorem: Theory, History, and Applications. Kaitlyn Shavelle, Princeton University. Grad Seminar (AST 542). Page 2. This was a talk on ...
  11. [11]
  12. [12]
    The Van Cittert Zernike Theorem
    Proof of the van Cittert-Zernike theorem can be found in a number of textbooks, eg. ``Optics'' by Born and Wolf, ``Statistical Optics'' by Goodman ...
  13. [13]
    [PDF] Chapter 9: Interference [version 1009.1.K] - Caltech PMA
    The two Fourier relations (9.7a), (9.7b) are called the van Cittert-Zernike Theorem. In. Ex. 9.7, we shall see that this theorem is a complex-variable version ...<|control11|><|separator|>
  14. [14]
    Van Cittert–Zernike Theorem, Spatial Coherence, and Scattering
    Feb 23, 2017 · The Fourier transform relationship provided by the van Cittert–Zernike theorem requires the source to be spatially incoherent. Spatial coherence ...
  15. [15]
    [PDF] 16.1 The van Cittert - Zernike theorem and radio interferometers
    The largest single-dish radio telescopes have diffraction-limited resolution no better than tens of arcseconds, but as we have noted, the vast majority of ...Missing: original paper
  16. [16]
    [PDF] Indirect Imaging 2 Two-Element Interferometer Van Cittert-Zernike ...
    van Cittert-Zernike theorem as Fourier pair. I(Ω) ⇔ ˜Γ(r/λ). Page 27. Overview with Equations (continued) van Cittert-Zernike theorem. ˜Γ(r) = Z Z source. I(Ω)e.
  17. [17]
    4.5 The Fourier Approximation & the van Cittert-Zernike theorem
    In this approximation, we assume that the different part of the sky are statistically independent to each other, meaning that the mutual coherency between two ...
  18. [18]
    On the diffraction theory of optical images - Journals
    On the diffraction theory of optical images. Harold Horace Hopkins. Google Scholar · Find this author on PubMed · Search for more papers by this author. Harold ...
  19. [19]
  20. [20]
  21. [21]
  22. [22]
  23. [23]
    [PDF] 2. Aperture Synthesis - NRAO
    Aperture Synthesis. The Complex Visibility. Page 9. • This FT relationship is the van Cittert-Zernike theorem, upon which synthesis imaging is based. • It means ...Missing: VLA | Show results with:VLA
  24. [24]
    [PDF] Fundamentals of astronomical optical Interferometry
    Van Cittert-Zernike theorem. Scientific motivations: – stellar interferometry: measuring stellar diameters. – ... Interferometry and angular resolution.
  25. [25]
    [PDF] Theoretical Study of Partially Coherent Imaging Systems using ...
    2.3 van Cittert–Zernike theorem . ... starts with the Fraunhofer diffraction formula and moves to the van Cittert–Zernike.
  26. [26]
    [PDF] Coherent Imaging with X-ray Free-Electron Lasers - JuSER
    partially-coherent beams. The van Cittert-Zernike theorem states that the mutual coherence function of a beam propagating from an incoherent source is given ...
  27. [27]
  28. [28]
  29. [29]
    [PDF] Principles of Adaptive Optics
    The Van Cittert–Zernike theorem (Roggemann and Welsh 1996) states that the spatial coherence proper- ties of an optical field are a Fourier transform of the ...
  30. [30]
    [PDF] ATMOSPHERIC WAVEFRONT SENSING AND CORRECTION ...
    Figure 4.2, Illustration of Van Cittert-Zemike theorem ... correction (adaptive ... In adaptive optics, understanding the nature of atmospheric turbulence is.
  31. [31]
    Speckle-Correlation Holographic Imaging: Advances, Techniques ...
    According to the van Cittert–Zernike theorem, the two-point intensity correlation at the far field is linked to the intensity distribution of the incoherent ...
  32. [32]
    [2202.07119] Multiphoton Quantum van Cittert-Zernike Theorem
    Feb 15, 2022 · Here, we introduce the quantum van Cittert-Zernike theorem to describe the scattering and interference effects of propagating multiphoton systems.Missing: microscopy imaging