Fact-checked by Grok 2 weeks ago

Huygens–Fresnel principle

The Huygens–Fresnel principle is a foundational concept in wave optics stating that every unobstructed point on a serves as a source of secondary spherical wavelets, each propagating outward with the same as the primary wave, and that the resulting wavefront at any subsequent point is formed by the superposition of these wavelets, accounting for their amplitudes and phases. Originally proposed by in 1690 as a qualitative method to explain wave propagation, reflection, and refraction using secondary wavelets, it was quantitatively refined by in 1815–1819, who incorporated Thomas Young's interference principle and introduced an obliquity factor to better model by suppressing backward-propagating wavelets. This synthesis provided a unified framework for understanding light's wave nature, particularly patterns observed in experiments like the double-slit setup, and was later rigorously derived by in 1882 using in scalar theory. In mathematical terms, the principle approximates the wave field \psi(\mathbf{r}) at a point as an integral over the : \psi(\mathbf{r}) = \int \frac{A}{r} e^{i k r} \psi(\mathbf{r}') d^2\mathbf{r}', where A incorporates the obliquity factor, r is the distance from aperture points \mathbf{r}' to the observation point, and k is the ; this holds under approximations such as far-field observation (z \gg aperture size) and small diffraction angles. The principle applies to scalar , including , , and water , enabling predictions of fringes and envelopes, as seen in Young's two-slit experiment where intensity I = I_1 + I_2 + 2\sqrt{I_1 I_2} \cos(\Delta\phi) arises from phase differences \Delta\phi. Beyond , the Huygens–Fresnel principle underpins modern applications in fields like acoustics for beam profiling, for diffraction quantification in signal propagation, and for , while recent geometrized formulations extend its utility to both near-field (Fresnel) and far-field ( without traditional approximations.

Historical Development

Origins in Huygens' Work

In 1678, composed his treatise Traité de la Lumière, which laid the groundwork for a wave theory of light by proposing that every point on a serves as a source of secondary spherical wavelets, and that the new is formed by the to these wavelets. This geometric construction allowed Huygens to model light propagation as a successive process through an elastic ether, where each wavelet expands spherically from its origin until its surface aligns with the overall advancing front. Although the work was not published until 1690, it provided a mechanistic explanation for light's finite speed, contrasting sharply with the instantaneous action implied by Isaac Newton's corpuscular theory in his (1704). Huygens' ideas drew partial inspiration from earlier wave-like conceptions of light, notably Robert Hooke's suggestion in (1665) that light consists of pulses propagating through the like sound waves. Building on this, Huygens applied his principle geometrically to derive the laws of , demonstrating that the incident and reflected rays make equal angles with by constructing the envelope of wavelets at a reflecting surface. This approach resolved longstanding puzzles about light's propagation, such as why it travels in straight lines despite its wave nature, by emphasizing the coherent forward advance of the . A key phenomenological assumption in Huygens' formulation was that secondary wavelets propagate only in the forward direction from the incident , without significant backward emission, which ensured straight-line propagation and avoided contradictions with observed motion. This rule, introduced without a detailed microscopic justification at the time, supported the wave model's viability against Newton's particle-based emissions. Later refinements, such as those by in the 19th century, would incorporate among these wavelets to explain and .

Fresnel's Contributions and Experimental Validation

In 1818, significantly advanced ' 1690 concept of secondary wavelets by incorporating Thomas Young's principle of , positing that diffraction patterns arise from the superposition of these wavelets with phase differences that lead to constructive and destructive . This addition addressed limitations in Huygens' purely geometric approach, enabling quantitative predictions of light intensity in diffracted fields through amplitude summation. Fresnel's work gained prominence amid a competition by the , which offered its for advancing understanding of to resolve debates between wave and corpuscular theories of . Despite initial skepticism from prominent corpuscular advocates like and on the judging committee, Fresnel submitted a comprehensive just before the deadline, earning the prize and marking a pivotal shift toward wave acceptance. A key validation came from the 1819 Poisson spot experiment, where Poisson—applying Fresnel's interference-based mathematics to a circular opaque disk—predicted an unexpected bright spot at the shadow's center, intending to discredit the wave model. , the committee president and a wave theory proponent, experimentally confirmed using a small metal disk and , demonstrating that wavelets from the disk's edge interfered constructively at the center, thus empirically supporting Fresnel's enhancements. To refine Huygens' assumption of uniformly forward-propagating wavelets, Fresnel introduced an obliquity factor in his formulation, conceptually weighting contributions from secondary sources based on their angular orientation relative to the , thereby improving agreement with observed patterns. This empirical adjustment resolved directional inconsistencies without full mathematical derivation at the time. In 1882, provided rigorous justification for the principle using and the scalar , deriving the obliquity factor and establishing its foundational role in wave propagation theory.

Core Concepts and Interpretation

Statement of the Principle

The Huygens–Fresnel principle asserts that every point on a given can be considered a source of secondary spherical wavelets that propagate forward from that point, with the subsequent wavefront formed as the to these expanding wavelets, and the at any location on the new wavefront resulting from the superposition and of all such wavelets. This conceptual framework, originally proposed by in 1690 and refined by in 1818, provides an intuitive means to visualize wave propagation without relying on particle-like rays. Intuitively, for a plane wave advancing through space, each infinitesimal segment of the initial wavefront emits a spherical wavelet that expands radially ahead, and after a short time interval, the crests of these wavelets align to form a parallel plane further along the propagation direction, maintaining the wave's shape and uniformity. This envelope construction highlights how the principle unifies diverse optical behaviors—such as straight-line propagation in uniform media—under a single wave-theoretic model, bridging macroscopic observations with the idea of distributed microscopic sources. Unlike ray optics, which traces light along straight paths ignoring wave nature, the Huygens–Fresnel approach inherently accounts for bending and spreading by treating waves as continuous disturbances rather than discrete trajectories, emphasizing interference as the key mechanism for amplitude variation. The principle's phenomenological character lies in its empirical foundation: it serves as a practical tool for predicting wave evolution rather than a from microscopic physics, effectively linking wavefronts to hypothetical secondary sources without specifying their origins. A crucial aspect is the restriction to forward-propagating wavelets, an rule introduced by Fresnel to suppress unphysical backward that would otherwise imply energy radiation in the opposite direction, consistent with experimental evidence that travels unidirectionally from its source. This forward bias ensures the model aligns with real-world observations, such as the absence of significant "wake" effects behind advancing .

Microscopic Model for Wave Propagation

The Huygens–Fresnel principle provides a semi-classical microscopic for wave propagation, envisioning each point on a as a "virtual source" that emits secondary spherical wavelets, much like the oscillatory disturbances of atoms or molecules in a propagating medium that collectively form the observable macroscopic wave envelope through superposition. This treats the not as a rigid entity but as a collection of emitters, each contributing to the forward advancement of the disturbance in a manner reminiscent of how microscopic vibrations in elastic media sustain longitudinal or transverse waves. By considering these point sources, the principle qualitatively explains how plane waves remain plane and spherical waves expand radially in free space, capturing the essence of wave without invoking detailed particle . Despite its intuitive appeal, the model harbors significant limitations that stem from its simplifying assumptions. Notably, it posits the emission of wavelets as effectively instantaneous from points on the initial , overlooking the finite speed at which disturbances propagate between "particles" in the medium, which leads to inaccuracies in describing the temporal evolution of short- or transient . Furthermore, the approach disregards fine-scale effects occurring on the order of the itself and depends on far-field approximations where observation distances greatly exceed the , rendering it unsuitable for near-field phenomena. For mathematical rigor, the principle is often anchored to Kirchhoff's boundary conditions, which apply to the wave equation to justify the selective forward propagation of wavelets while suppressing backward components. In homogeneous media devoid of absorption, the Huygens–Fresnel model effectively simulates unobstructed wave propagation by assuming constant wave speed and no dissipative losses, allowing the envelope of interfering wavelets to faithfully reconstruct the advancing wavefront. This setup aligns well with scenarios in classical optics, such as light transmission through uniform glass or air, where the principle's wavelet superposition yields the correct phase and amplitude progression. Conceptually, it approximates solutions to the Helmholtz equation, the time-independent form of the scalar wave equation that describes harmonic waves in lossless, isotropic environments, providing a practical heuristic for predicting field distributions without solving the full differential equation directly. While invaluable for pedagogical and qualitative insights into wave behavior, the Huygens–Fresnel principle remains a phenomenological construct rather than a genuine microscopic theory, as it predates and does not emerge directly from that govern electromagnetic interactions at the fundamental level; contemporary theoretical extensions, including vectorial and quantum formulations, address these gaps in later discussions.

Mathematical Formulation

Classical Expression and Derivation

The classical expression of the Huygens–Fresnel principle provides a mathematical framework for calculating the wave field at an observation point P based on the field distribution over a wavefront or aperture surface S. For a monochromatic scalar wave field U(Q) at points Q on S, the field U(P) at P is given by the integral U(P) = \frac{1}{i\lambda} \int_S U(Q) \frac{e^{iks}}{s} \frac{1 + \cos\theta}{2} \, dS, where s is the distance from Q to P, k = 2\pi / \lambda is the wavenumber with wavelength \lambda, and \theta is the angle between the normal to the surface at Q and the line connecting Q to P. This formula represents the superposition of secondary spherical wavelets emanating from each point on S, weighted by the obliquity factor \frac{1 + \cos\theta}{2} to account for directional emission. The derivation of this expression stems from a rigorous application of Green's theorem to the scalar wave equation. Consider the Helmholtz equation \nabla^2 U + k^2 U = 0 satisfied by the monochromatic field U in free space. By applying Green's second identity to U and the Green's function G(Q, P) = \frac{e^{iks}}{4\pi s} (the fundamental solution to the Helmholtz equation), the field at P inside a volume V bounded by surface \Sigma is expressed as a surface integral over \Sigma: U(P) = \oint_\Sigma \left( U(Q) \frac{\partial G}{\partial n_Q} - G \frac{\partial U}{\partial n_Q} \right) dS, where \partial / \partial n_Q denotes the normal derivative at Q. For diffraction problems, the surface \Sigma encloses the aperture S and a large hemispherical boundary at infinity, where the field and its derivative vanish due to radiation conditions; the integral over the infinite boundary contributes negligibly, reducing the expression to an integral over the aperture S. The Kirchhoff boundary condition assumes U and \partial U / \partial n on S are known from the incident field, and for normal incidence of a plane wave, \partial U / \partial n \approx i k U, with the obliquity factor emerging from the combination of the \partial G / \partial n_Q term (which includes \cos\theta) and the boundary conditions, leading to the Huygens–Fresnel form after simplification. This derivation, formalized by Kirchhoff in 1882, provides a boundary-value solution to the wave equation that justifies the heuristic Huygens–Fresnel construction. Key assumptions underpin this formulation: the waves are monochromatic, ensuring time-harmonic dependence and the validity of the ; the scalar treats the field as a scalar , suitable for or when effects are negligible; and the far-field (Fraunhofer) holds when the observation distance is much larger than the size and (s \gg a^2 / \lambda, where a is the dimension), allowing neglect of higher-order variations. The principle's forward-only propagation, avoiding unphysical backward waves, is justified by the use of the retarded , which incorporates through the time-delayed e^{iks} derived from the wave equation's initial-value problem. This ensures that contributions to U(P) arise solely from sources on the ahead of P, consistent with the retarded potentials in electromagnetic , while the obliquity factor further suppresses backward components (\theta \approx \pi).

Obliquity Factor and Key Approximations

The obliquity factor represents a key refinement introduced by Fresnel to the original Huygens principle, which assumed isotropic emission of secondary from each point on a . This correction accounts for the predominantly forward-directed nature of wave propagation by modulating the amplitude of contributions based on the direction relative to the wavelet normal, thereby suppressing backward and grazing-angle emissions that would otherwise contradict observed . The factor is mathematically expressed as K(\chi) = \frac{1 + \cos \chi}{2}, where \chi is the angle between the normal to the secondary source on the and the line connecting it to the observation point. This form ensures full (K=1) for normal incidence (\chi = 0^\circ) and reduces it to half (K=0.5) at angles (\chi = 90^\circ), demonstrating a significant —such as a 50% drop—for oblique directions that helps align the model with empirical patterns. Although Fresnel's introduction of the obliquity factor was empirical, later provided a rigorous foundation in by deriving it from the using and boundary conditions on the , confirming its necessity and resolving the ad-hoc aspects of the earlier formulation while preserving consistency with electromagnetic . In practical applications of the Huygens–Fresnel integral, several approximations simplify computations while maintaining accuracy for specific regimes. The applies to near-field , retaining quadratic terms in the phase expansion to capture curved propagation over distances where the observation point is not infinitely far, typically when the Fresnel number F = a^2 / (\lambda z) \gtrsim 1 (with a as size, \lambda , and z distance)./06%3A_Scalar_diffraction_optics/6.07%3A_Fresnel_and_Fraunhofer_Approximations) In contrast, the Fraunhofer approximation is valid in the far field (F \ll 1), neglecting higher-order terms to assume plane-wave incidence and yielding a pattern as the of the aperture function, independent of exact distance beyond the far-field condition./06%3A_Scalar_diffraction_optics/6.07%3A_Fresnel_and_Fraunhofer_Approximations) The paraxial approximation further streamlines these by assuming small propagation angles (\theta \ll 1), allowing \cos \chi \approx 1 - \chi^2/2 and neglecting transverse variations, which is essential for many optical systems but limits validity to near-axis regions./06%3A_Scalar_diffraction_optics/6.07%3A_Fresnel_and_Fraunhofer_Approximations) While the standard Huygens–Fresnel formulation is scalar and treats as a monochromatic wave without , vector extensions incorporate electromagnetic boundary conditions to address polarization-dependent effects, such as differing for s- and p-polarized , though these increase .

Applications in Classical Optics

Refraction

The Huygens–Fresnel principle provides a wave-based explanation for at the interface between two media, where the changes from v_1 in the first medium to v_2 in the second, with v_2 < v_1 for denser media. Consider a plane wave incident obliquely on a planar boundary; each point on the incident wavefront acts as a source of secondary wavelets that propagate at the respective speeds in each medium. The new wavefront in the second medium forms as the envelope tangent to these wavelets, resulting in a bending of the wavefront toward the normal due to the slower propagation speed. In the geometric construction, imagine the incident wavefront AB reaching the boundary at point A first, while point B reaches later. From A, a wavelet expands into the second medium at speed v_2 over time t, covering distance v_2 t. Meanwhile, from B, the wavelet in the first medium covers v_1 t along the boundary direction. The refracted wavefront is the line connecting the envelope points, forming two right triangles sharing the hypotenuse along the boundary. The angles of incidence i and refraction r satisfy \sin i / v_1 = \sin r / v_2, leading to \sin i / \sin r = v_1 / v_2. Since refractive indices are n_1 = c / v_1 and n_2 = c / v_2, this yields Snell's law: n_1 \sin i = n_2 \sin r. Phase matching across the boundary ensures continuity of the wavefront, as the optical path lengths align to maintain coherent propagation. For a specific example, light transitioning from air (n_1 \approx 1) to glass (n_2 \approx 1.5) bends the wavefront such that the angle of refraction is smaller than the incidence angle, without considering reflection in the pure refraction case; the slower speed in glass causes wavelets to lag, tilting the envelope. Unlike ray optics, which treats light as straight paths obeying empirical rules, the Huygens–Fresnel approach highlights the wave nature by resolving setups like total internal reflection, where beyond the critical angle, no real refracted wavefront forms due to the impossibility of \sin r > 1, instead leading to evanescent waves parallel to the boundary.

Diffraction

The Huygens–Fresnel principle explains diffraction as the interference of secondary wavelets emanating from points along a wavefront obstructed by an aperture or edge, leading to the bending and spreading of light beyond geometric predictions. In a basic single-slit diffraction setup, each point across the open slit acts as a source of spherical wavelets that propagate forward and interfere at a distant screen, producing a central bright maximum flanked by alternating dark minima and secondary maxima due to path length differences. For instance, the first minimum occurs where wavelets from the slit edges are out of phase by π radians with those from the center, resulting in destructive interference. A striking demonstration is the Poisson spot, or Arago spot, observed in the shadow of a circular obstacle, where the principle predicts a bright spot at the center due to constructive interference of wavelets from the obstacle's edge, countering the expectation of uniform darkness from ray optics. This phenomenon arises because wavelets from symmetrically opposite points on the circumference travel equal distances to the shadow's center, reinforcing in phase. Key to these patterns are phase differences among wavelets, which cause constructive in bright fringes and destructive in dark regions, qualitatively predicting the angular spread of fringes inversely proportional to size. The principle refutes straight-ray propagation by showing how wavelets from edge points bend around corners or obstacles, enabling illumination in shadowed areas. In near-field diffraction, Fresnel zones divide the into concentric regions where each zone contributes differing in by π, with odd zones adding constructively and even zones destructively at the observation point, allowing qualitative assessment of variations close to the . The obliquity factor modulates amplitudes based on emission , reducing contributions from oblique directions to better match observations. Modern applications include computational diffraction simulations that discretize apertures into wavelet sources and sum their fields numerically, aiding design for lenses and holograms by predicting patterns efficiently.

Generalizations and Extensions

Generalized Huygens' Principle

The generalized Huygens' principle extends the classical optical formulation to the propagation of arbitrary solutions to the wave equation, representing the wave field at a point as a over secondary sources on a , weighted by the appropriate . This approach provides an exact mathematical framework for wave evolution, applicable to , , and tensor fields satisfying partial equations. For a \psi obeying the wave equation \left( \nabla^2 - \frac{1}{c^2} \frac{\partial^2}{\partial t^2} \right) \psi = 0, the field at position \mathbf{x}' and time t' can be expressed using the retarded G(\mathbf{x}', t'; \mathbf{x}, t), which satisfies the same equation with a delta-function source and enforces by vanishing for t' < t. In the volume integral form for initial data, this yields \psi(\mathbf{x}, t) = \int \left[ G(\mathbf{x}, t; \mathbf{x}', 0) \frac{\partial \psi}{\partial t}(\mathbf{x}', 0) - \psi(\mathbf{x}', 0) \frac{\partial G}{\partial t'}(\mathbf{x}, t; \mathbf{x}', 0) \right] d^3\mathbf{x}' over the initial , though over arbitrary closed surfaces are equivalent via for source-free regions. This generalization applies directly to matter waves as proposed by de Broglie, where particles exhibit wave-like propagation governed by the , which is first-order in time and thus inherently forward-propagating without backward tails, aligning with the strict Huygens condition. Similarly, for electromagnetic fields, the principle manifests through vector potentials satisfying the in Lorentz , allowing exact reconstruction of fields from boundary values using dyadic s that incorporate the nature of the sources. In inhomogeneous , where the wave speed varies spatially, the approach yields exact solutions by solving the modified , enabling propagation through gradients without approximation. A key feature is Hadamard's minimal condition for the strict Huygens principle, which requires that wave disturbances propagate without "tails"—residual fields lingering behind the —achieved when the fundamental solution of the wave operator has support only on the , as in three spatial dimensions for massless scalar fields. This condition ensures sharp , free of tails, and holds for the d'Alembertian operator in odd-dimensional . Applications extend to acoustics, where sound waves in fluids obey the scalar , and the principle models around obstacles using surface integrals of and potentials. In , it underpins the causal evolution of wave functions, linking classical wave to probabilistic interpretations via retarded propagators that respect the no-signaling . Forward is rigorously enforced by the retarded , which incorporates through the theta-function \theta(t' - t - |\mathbf{x}' - \mathbf{x}|/c), preventing influences from future or spacelike-separated points.

Behavior in Other Spatial Dimensions

The Huygens–Fresnel principle, which posits that every point on a wavefront acts as a source of secondary wavelets leading to sharp propagation, holds strictly only in odd spatial dimensions, such as one and three dimensions, but fails in even spatial dimensions by producing lingering "tails" in the wave disturbance. This key observation was first made by Jacques Hadamard in his analysis of the wave equation around 1900, highlighting a fundamental dimensional dependency in wave propagation. The mathematical basis for this behavior lies in the solutions to the wave equation in n spatial dimensions, where sharp propagation—characteristic of Huygens' principle—occurs exclusively for odd n due to the properties of spherical waves in those spaces. In odd dimensions greater than or equal to three, such as five or seven, the fundamental solution () is supported solely on the , ensuring that disturbances propagate without diffusion or tails inside the . Conversely, in even dimensions, the exhibits support both on and inside the , resulting in tails that represent slower-propagating components trailing the main . A representative example is , where waves from a expand as cylindrical rather than spherical waves, leading to persistent tails that cause the disturbance at a point to depend on initial conditions throughout the entire disk behind the , rather than just its . This contrasts with the clean, boundary-only dependence in three dimensions and underscores why the Huygens–Fresnel principle requires modification, such as through representations accounting for the interior contributions, in even-dimensional settings. This dimensional variation has implications for , particularly in models involving curved spacetimes or , where wave propagation can exhibit tail-like behaviors if the effective spatial dimensionality is even, influencing phenomena like gravitational radiation in higher-dimensional theories. In contexts such as or braneworld cosmology, understanding these tails is crucial for predicting signal propagation in extra-dimensional geometries, though such applications remain largely theoretical.

Modern Theoretical Perspectives

Feynman's Path Integral Formulation

The Huygens–Fresnel principle, which describes wave propagation as the superposition of secondary wavelets from points on a wavefront, finds a quantum mechanical analog in Richard Feynman's path integral formulation, where each wavelet corresponds to a classical path contributing to the total amplitude. In this approach, the probability amplitude for a photon or particle to propagate from an initial point to a final point is given by the sum over all possible paths, each weighted by a phase factor e^{iS/\hbar}, where S is the classical action along that path and \hbar is the reduced Planck's constant. This summation naturally incorporates interference effects, mirroring the constructive and destructive superposition of wavelets in the classical principle. Feynman's formulation, developed in the late 1940s as part of his work on (), provides a probabilistic of wave phenomena. For instance, in the with photons, the sums contributions from all trajectories passing through either slit, leading to patterns where paths with similar phases add constructively and those with differing phases cancel, explaining the observed intensity distribution on the screen. The modern photon \psi, derived from this formalism, yields the intensity as |\psi|^2, which aligns with the envelope of probabilities in the Huygens–Fresnel description, ensuring consistency between quantum and classical in the high-photon-number limit. This approach extends de Broglie's hypothesis of matter waves, applying the summation over paths to massive particles like electrons, where the associated wave propagates according to Huygens' principle with wavelength \lambda = h/p (Planck's constant h over p). In the , computational implementations of path integrals have advanced simulations of complex optical systems, such as linear in multi-mode interferometers, using efficient path-sum algorithms for low-depth circuits that offer advantages in runtime and memory efficiency while capturing interference effects.

Connections to Quantum Field Theory

In quantum field theory (QFT), the Huygens–Fresnel principle provides a classical visualization of field , where every point on a acts as a source of secondary wavelets analogous to the propagation of quantum fields through Green's functions. For free fields in homogeneous , these Green's functions satisfy the Huygens–Fresnel principle, representing the response to an impulse and enabling the superposition of solutions to the wave equation that underpin field evolution. This connection arises because the in QFT, which is the time-ordered of field operators, mirrors the spherical wavelet emission in the principle, ensuring and support on the in three spatial dimensions. This adherence to the is tied to the principle's validity in odd spatial dimensions, such as three, where wave propagation is sharp without tails, unlike in even dimensions. The propagation of in QFT aligns with this framework, as these operators evolve via the same Green's functions that embody the Huygens–Fresnel superposition. Path integrals over field configurations in QFT yield amplitudes that incorporate effects consistent with wavelet superposition, extending the principle to quantized fields where all possible paths contribute democratically to the amplitude. In homogeneous , Lorentz invariance preserves the principle's strict adherence to the , preventing signals from propagating inside it and maintaining the relativistic structure of free-field theories. This formulation connects to theory, where propagators mediate interactions, and the resulting amplitudes reflect the of wavelets in processes, providing a bridge between classical wave and quantum . In modern applications, such as simulations, the principle informs models of propagation and at the quantum level, ensuring consistency between classical and quantum descriptions.

References

  1. [1]
    8.5: The Huygens Principle - Physics LibreTexts
    Mar 5, 2022 · Deriving the Huygens principle. However, another approximation, called the Huygens (or “Huygens-Fresnel”) principle, is very instrumental ...
  2. [2]
    Geometrization of the Huygens–Fresnel principle - AIP Publishing
    May 17, 2024 · A modern restatement of the Huygens–Fresnel principle is as follows: “Every unobstructed point of a wavefront serves as a source of secondary ...A brief history of the Huygens... · A contemporary advance on... · Salient predictions
  3. [3]
    Huygens Principle - an overview | ScienceDirect Topics
    Huygens' principle is defined as a method for understanding wave propagation, stating that each point on a wave front acts as a source of secondary wavelets ...
  4. [4]
    [PDF] TREATISE ON LIGHT - Thomas Aquinas College
    wrote this Treatise during my sojourn in France twelve years ago, and I communicated it in the year 1678 to the learned persons who then.Missing: Traité Lumière
  5. [5]
    huygens' traite de la lumiere and newton's opticks - jstor
    Although Huygens died in 1695, we will be able to confidently predict what his reaction to Book II of the. Opticks (1704) would have been once we understand his ...
  6. [6]
    [PDF] Christiaan Huygens' Wave Theory of Light: A Major Contribution to ...
    Explaining optical phenomena strictly in terms of matter, motion, and impact, Huygens postulated that light was caused by a series of spherical waves being ...
  7. [7]
    [PDF] Learning Optics from the History - Repositorio Institucional UCA
    Nov 28, 2019 · In 1678, Huygens published his “Traité de la Lumiere”, where he enunciated the principle according to which every point the “aether” upon ...
  8. [8]
    Augustin Fresnel and the wave theory of light - Photoniques
    Fresnel pursued his experiments between 1814 and 1818, introducing the notion of wavelengths, exploiting the Huygens principle, and formulating the mathematical ...
  9. [9]
    July 1816: Fresnel's Evidence for the Wave Theory of Light
    Fresnel, Augustin. (1818) “Memoir on the Diffraction of Light,” The Wave Theory of Light: Memoirs by Huygens, Young and Fresnel. Woodstock, GA: ...
  10. [10]
    Reproducing the Fresnel-Arago experiment to illustrate physical optics
    Fresnel, together with the committee president,. François Arago, put Poisson's prediction to the test, and Arago announced to the world that there is indeed a ...
  11. [11]
  12. [12]
    Huygens' Principle geometric derivation and elimination of the wake ...
    Oct 12, 2021 · Huygens' Principle (1678) implies that every point on a wave front serves as a source of secondary wavelets, and the new wave front is the tangential surface ...
  13. [13]
    [PDF] Chapter 15S Fresnel-Kirchhoff diffraction - bingweb
    Jan 22, 2011 · Fresnel was able to account for diffraction by supplementing Huygen's construction with the postulate that the secondary wavelets mutually ...<|separator|>
  14. [14]
    Huygens–Fresnel Principle
    $$K(\theta)$ is known as the obliquity factor. Previously (in Section 10.9) ... diffraction, whereas near-field diffraction is termed Fresnel diffraction.
  15. [15]
    [PDF] Sur la diffraction de la lumière - Mémoire d'Augustin Fresnel ...
    F. SUR LA DIFFRACTION DE LA LUMIÈRE,. 343 attribuer cette diversité de couleurs ou de sensations produites sur l'organe de la vue à des différences de masse ou ...
  16. [16]
    Vector diffraction based on 3D Huygens-Fresnel principle and vector ...
    We proposed vector diffraction equations based on 3D Huygens-Fresnel principle and vector decomposition in Cartesian coordinate.
  17. [17]
  18. [18]
    [PDF] Physics for Scientists & Engineers 2
    Apr 8, 2005 · Derivation of Snell's Law. ▫ Now let's use a Huygens' construction to derive Snell's Law for refraction between two optical media with different.<|control11|><|separator|>
  19. [19]
    [PDF] Physics II for Science and Engineering
    Law of Refraction (Snell's Law). • Like the law of reflection, the law of ... Derivation of Snell's Law Using Huygens' Principle sin. / sin . Also, sin.
  20. [20]
    Single slit diffraction
    Huygens' principle tells us that each part of the slit can be thought of as an emitter of waves. All these waves interfere to produce the diffraction pattern.
  21. [21]
    [PDF] Single-slit Diffraction - SMU Physics
    According to Huygens's principle, each portion of the slit acts as a source of light waves. Hence, light from one portion of the slit can interfere with.
  22. [22]
    84.12 -- Laser beam diffracted by a ball bearing (Arago's spot)
    Poisson noted that Fresnel's theory predicted that there should be a bright spot in the center of the shadow of a circular object (which Fresnel had not done).
  23. [23]
    Why is there no Poisson spot in a solar eclipse? - AIP Publishing
    Dec 1, 2024 · The Poisson spot is an optical phenomenon in which a bright spot of light is seen at the center of a shadow caused by a spherical object.
  24. [24]
    [PDF] Diffraction
    Such a ring or circle is called a Fresnel zone. If we increase h further, such that ∆ = λ, we get a ring of sources that produce fields in phase with the. LOS ...
  25. [25]
    [PDF] Fresnel Diffraction - Waves & Oscillations
    Huygens-Fresnel Principle. Huygens: source of spherical wave. Problem: it must also go backwards, but that is not observed in experiment inclination factor.<|separator|>
  26. [26]
    Simulating multiple diffraction in imaging systems using a path ...
    We present a method for simulating multiple diffraction in imaging systems based on the Huygens–Fresnel principle. The method accounts for the effects of ...
  27. [27]
    [PDF] Huygens' principle and Hadamard's conjecture
    If for a Huygens operator L the function f vanishes outside a neighbourhood of x 0 ~ I~, then u§ van- ishes outside a neighbourhood of C+ (x0) in accordance.
  28. [28]
    Huygens-Fresnel Acoustic Interference and the Development of ...
    Apr 12, 2017 · A single propagating surface acoustic wave interacts with the proximal channel wall to produce a knife-edge effect according to the Huygens-Fresnel principle.
  29. [29]
    Retarded Green's functions in perturbed spacetimes for cosmology ...
    Dec 9, 2011 · This paper is primarily concerned with how to solve for the retarded Green's functions of the minimally coupled massless scalar φ , photon A μ ...
  30. [30]
    Huygens' Principle - MathPages
    The Huygens-Fresnel Principle is adequate to account for a wide range of optical phenomena, and it was later shown by Gustav Kirchoff (1824-1887) how this ...Missing: limitations Kirchhoff
  31. [31]
    [PDF] 7 Wave Equation in Higher Dimensions
    In fact, as we will see, for n ≥ 3 and odd, Huygens's principle holds. That is, all signals travel at exactly speed c. In even dimensions, however, that is not ...
  32. [32]
    Anomalously small wave tails in higher dimensions | Phys. Rev. D
    We show that in six and higher even dimensions there exist exceptional potentials for which the tail has an anomalously small amplitude and fast decay. Tails ...Missing: extra | Show results with:extra
  33. [33]
    [2105.00948] Path Integrals: From Quantum Mechanics to Photonics
    May 3, 2021 · View a PDF of the paper titled Path Integrals: From Quantum Mechanics to Photonics, by Charles W. Robson and 3 other authors. View PDF.
  34. [34]
    Huygens-Fresnel principle: Analyzing consistency at the photon level
    Jun 22, 2018 · View a PDF of the paper titled Huygens-Fresnel principle: Analyzing consistency at the photon level, by Elkin A. Santos and 1 other authors.
  35. [35]
    Feynman path sum approach for simulation of linear optics - arXiv
    Oct 30, 2025 · The Feynman path integral formalism has inspired the development of memory-efficient and parallelizable classical algorithms for simulating ...Missing: computational 2020-2025
  36. [36]
  37. [37]
    Huygens-Fresnel principle: Analyzing consistency at the photon level
    Apr 23, 2018 · Here we show that because of the close relation existing between the FRFT and the Fresnel diffraction integral, this propagator can be written ...<|control11|><|separator|>
  38. [38]
    Huygens' principle, the free Schrodinger particle and the quantum ...
    Aug 17, 2001 · Huygens' principle following from the d'Alembert wave equation is not valid in two-dimensional space. A Schrodinger particle of vanishing ...
  39. [39]
    Why is Huygens' principle only valid in an odd number of spatial ...
    Aug 3, 2014 · Hadamard essentially took a slice through a cylindrical wave in three dimensions to get a circular wave in two dimensions (descending one ...How does Huygens's Principle Work? - Physics Stack ExchangeNature of Huygens' principle - Physics Stack ExchangeMore results from physics.stackexchange.com
  40. [40]
    Toward a localized -matrix theory | Phys. Rev. D
    This suppression is achieved without imposing the restrictions owing to the Huygens-Fresnel's principle, but it is rather a consequence of the inherent boundary ...
  41. [41]
    Reliable, efficient, and scalable photonic inverse design empowered ...
    Jan 27, 2025 · For Huygens–Fresnel Stitch, this method is inspired by the Huygens–Fresnel principle, the fundamental principle of optical field propagation.
  42. [42]
    Quantum wave simulation with sources and loss functions
    Sep 5, 2025 · We present a quantum algorithmic framework for simulating linear, anti-Hermitian (lossless) wave equations in heterogeneous, anisotropic, ...Missing: Huygens- Fresnel