Fact-checked by Grok 2 weeks ago

Eikonal equation

The eikonal equation is a nonlinear that models the high-frequency limit of wave , approximating the function of waves in media where the is much smaller than variations in the speed. It takes the general form |\nabla u(\mathbf{x})| = n(\mathbf{x}), where u is the eikonal function representing the or , and n(\mathbf{x}) is the or inverse speed function in the medium. This equation captures the of paths, ensuring that solutions align with of least time for wave travel. Originating from geometrical optics, the eikonal equation derives from the under the short-wavelength approximation, where the wave amplitude is assumed slowly varying compared to the rapidly oscillating . The term "eikonal" stems from word εἰκών (eikōn), meaning "" or "," reflecting its roots in optical imaging and ray tracing. It is closely tied to , which states that light rays follow paths minimizing travel time, and the eikonal equation serves as its differential formulation, enabling derivations of phenomena like at refractive interfaces. In homogeneous media, solutions yield straight-line rays, while in inhomogeneous cases, rays curve according to variations in n(\mathbf{x}). Solutions to the eikonal equation are often viscosity solutions, a that handles discontinuities and ensures uniqueness for boundary value problems, such as the distance function in satisfying |\nabla d| = 1 with d = 0 on the boundary. Numerically, it is solved using methods like fast marching or sweeping algorithms, which propagate along characteristics in O(N \log N) time for N grid points, making it practical for large-scale computations. Beyond , the eikonal equation finds applications in for travel-time , computer vision for shape-from-shading and segmentation, for and MRI reconstruction, and for front-tracking. In heterogeneous media, it models effective distances weighted by local speeds, aiding homogenization theory for wave propagation in composites. Its hyperbolic nature and connection to Hamilton-Jacobi equations underscore its role in and variational problems across physics and .

Mathematical Foundations

Definition and Formulation

The eikonal equation is a first-order central to the high-frequency approximation of propagation problems. Its standard formulation is given by |\nabla T(\mathbf{x})| = n(\mathbf{x}), where T(\mathbf{x}) denotes the or travel time at \mathbf{x} \in \mathbb{R}^d, \nabla T is the of T, and n(\mathbf{x}) represents the or slowness, defined as the of the local speed. This equation is typically supplemented with boundary conditions, such as T = 0 on an initial surface or curve \Gamma, which specifies the starting point for propagation; here, T(\mathbf{x}) interprets as the optical path length or the arrival time of the wavefront at \mathbf{x}. Variations of the equation account for different medium properties. In the isotropic case with constant speed, n(\mathbf{x}) is uniform, yielding solutions where T scales linearly with from the . In the inhomogeneous case with position-dependent speed, n(\mathbf{x}) varies spatially, enabling modeling of heterogeneous media where propagation speed changes with location. Geometrically, the level sets of T(\mathbf{x}) = t for constant t describe wavefronts that advance normal to themselves at speed $1/n(\mathbf{x}).

Derivation from Wave Propagation

The eikonal equation emerges as the leading-order approximation in the high-frequency of the scalar , which describes time-harmonic wave propagation in inhomogeneous media. The Helmholtz equation takes the form \nabla^2 u + k^2 n^2(\mathbf{x}) u = 0, where u(\mathbf{x}) is the complex-valued wave amplitude, k = \omega / c_0 is the reference wavenumber with angular frequency \omega and speed c_0, and n(\mathbf{x}) is the position-dependent refractive index. In the Wentzel-Kramers-Brillouin (WKB) approximation, valid for large k \to \infty, the solution is postulated as u(\mathbf{x}) \approx A(\mathbf{x}) \exp(i k S(\mathbf{x})), separating the slowly varying amplitude A from the rapidly oscillating phase S. Substituting this ansatz into the Helmholtz equation, collecting terms by powers of k, and retaining only the dominant O(k^2) contribution yields the eikonal equation |\nabla S|^2 = n^2(\mathbf{x}), with subsequent orders providing a transport equation $2 \nabla S \cdot \nabla A + (\Delta S) A = 0 for the amplitude. Here, S(\mathbf{x}) represents the optical path length, and in the section's notation, the eikonal function T coincides with S (or travel time in units where c_0 = 1). The approximation holds by neglecting lower-order terms involving second derivatives of A and \Delta S, which become insignificant when the wavelength $2\pi / k is much smaller than the scale of medium variations. This derivation extends naturally to the time-dependent scalar \frac{\partial^2 u}{\partial t^2} = c^2(\mathbf{x}) \nabla^2 u, governing wave propagation with local speed c(\mathbf{x}). For high-frequency solutions, assume u(\mathbf{x}, t) \approx A(\mathbf{x}, t) \exp(i \omega (t - \tau(\mathbf{x}))), where \omega \gg 1 and \tau is the . Inserting this form and expanding asymptotically in powers of $1/\omega produces, at leading order, the eikonal equation |\nabla \tau|^2 = \frac{1}{c^2(\mathbf{x})}, with \tau(\mathbf{x}) interpreting as the travel time along characteristics. The next-order terms yield an amplitude transport equation analogous to the Helmholtz case, ensuring conservation of wave energy flux along rays. This asymptotic approach underscores the eikonal equation's connection to geometric optics, where waves propagate along rays perpendicular to wavefronts defined by constant phase.

Historical Development

Origins in Optics

The origins of the eikonal equation trace back to the with Pierre de Fermat's principle of least time, which posits that light travels between two points along the path that minimizes the travel time. This principle provided a variational foundation for understanding light propagation in inhomogeneous media, serving as a precursor to the eikonal's role in describing optical paths. Fermat formulated this principle in the 1650s, using it to derive the law of independently of earlier empirical observations. Building on this, had earlier proposed a related formulation in the 1630s for the laws of and , though his approach assumed varying speeds in media that contrasted with Fermat's temporal minimization. These 17th-century developments established the conceptual basis for paths as extrema of an optical length , central to the eikonal framework. In the late , the term "eikonal" was introduced by Heinrich Bruns in 1895, derived from word for "image," to denote a in variational that maps wavefronts to ray paths. Bruns' work formalized the eikonal as a tool for analyzing light ray bundles in complex optical systems, emphasizing its geometrical interpretation without reliance on wave theory. Early 19th-century advancements by Augustin-Jean Fresnel further connected these ideas to wavefront propagation, integrating ray optics with emerging wave descriptions of light to explain how rays remain tangent to advancing wavefronts in refracting media. Fresnel's contributions, particularly through his elastic theory of light around 1818, highlighted the continuity between geometrical rays and wave surfaces, laying groundwork for the eikonal's interpretation as the phase function in high-frequency approximations.

Evolution in Hamiltonian Mechanics

The mathematical formalization of the eikonal equation advanced significantly in the 19th century through Rowan Hamilton's development of characteristic functions in during the 1830s. Hamilton introduced the principal characteristic function, often denoted as V, which represents the between two points along a ray in a medium with varying . This function satisfies a that is precisely the eikonal equation, |\nabla V| = n(\mathbf{r}), where n is the refractive index. Hamilton's framework bridged and by deriving for rays analogous to those in mechanics, with the eikonal serving as the for ray trajectories. His approach emphasized the underlying ray paths, minimizing the , and laid the groundwork for treating propagation as a . Building on Hamilton's ideas, provided a key in the 1840s through his work on variational calculus and . In his lectures on delivered in 1842–1843 and later published, Jacobi extended the Hamilton-Jacobi equation to systems with multiple , introducing the as a in the of constrained variational problems. This transformed the eikonal into a more versatile tool within the broader framework of , where the principal function—equivalent to the eikonal—facilitates the and solution of complex ray paths. Jacobi's contributions clarified the , emphasizing the equation's role in optimizing paths via least action principles, and connected optical phenomena to general mechanical systems without relying on specific optical assumptions. The eikonal equation's scope expanded into relativistic contexts with David Hilbert's 1915 formulation of the foundations of physics, where variational methods yielded the of . In this framework, the eikonal describes the propagation of as null in curved , with the corresponding to the invariant interval along rays satisfying ds^2 = 0. Hilbert's axiomatic approach integrated electromagnetic and gravitational fields variationally, implicitly incorporating the eikonal as the phase function for high-frequency wave solutions in gravitational fields, thus linking to the geodesic motion of photons. This connection highlighted the equation's universality, treating rays as extremals in metrics. Twentieth-century refinements synthesized these developments in authoritative treatments of . Notably, and Emil Wolf's 1959 textbook provided a comprehensive analysis, deriving the eikonal equation from the scalar in the short-wavelength limit and exploring its implications for tracing, evolution, and . Their work emphasized practical computations of paths in inhomogeneous media and aberration theory, attributing the eikonal's foundational role to Hamilton's characteristic functions while extending applications to modern optical systems. This synthesis bridged 19th-century variational insights with wave-based derivations, solidifying the eikonal's place in both theoretical and .

Physical Interpretations

Geometrical Optics and Ray Tracing

In , the eikonal equation governs the propagation of light rays in inhomogeneous media by describing the evolution of , which are level sets of the eikonal function T(\mathbf{x}), the from a reference . Rays represent the orthogonal trajectories to these level sets, indicating the direction of energy flow perpendicular to the . The path of a parameterized by s follows the \frac{d\mathbf{x}}{ds} = \frac{\nabla T}{|\nabla T|}, where \nabla T points normal to the and |\nabla T| = n(\mathbf{x}), with n(\mathbf{x}) denoting the . This formulation arises from , which posits that rays minimize the \int n \, ds, ensuring stationary travel time along the . A key consequence of the eikonal equation is of at between media with different refractive indices. Consider a planar where n jumps from n_1 to n_2; the eikonal T remains continuous across the boundary, while its tangential component is preserved to maintain phase matching. For incident and refracted rays with angles \theta_1 and \theta_2 relative to , the condition n_1 \sin \theta_1 = n_2 \sin \theta_2 emerges directly from the constancy of the tangential slowness p = \frac{\sin \theta}{c} = \frac{n \sin \theta}{c_0}, where c_0 is the speed in . This ensures the ray bends such that the is , validating the eikonal approximation for high-frequency waves. The eikonal framework breaks down at caustics, regions where rays converge and intersect, forming singularities in the ray density. These occur when the Jacobian of the ray mapping vanishes, leading to focusing points or lines where the amplitude diverges as $1/\sqrt{\delta} near the caustic, with \delta the distance from it. Such singularities, exemplified by fold or cusp caustics in lens systems, signal the limits of geometrical optics, as diffraction effects from the full wave equation become prominent beyond the high-frequency approximation. The eikonal equation aligns with Huygens' principle in the geometrical optics limit, where wavefronts advance as the envelope of secondary wavelets emanating from points on the prior wavefront, each expanding at local speed c = 1/n. Rays, as normals to these evolving level sets of T, trace the direction of wavelet propagation, capturing the normal advancement without diffraction in homogeneous or slowly varying media. This geometric interpretation unifies ray and wavefront descriptions for light propagation.

Continuous Shortest-Path Formulations

The eikonal equation arises as a Hamilton-Jacobi equation characterizing the function in Riemannian or Finsler geometry, providing a continuous of shortest-path problems beyond wave contexts. Specifically, the T(\mathbf{x}) to the eikonal equation |\nabla T(\mathbf{x})| = n(\mathbf{x}) with boundary condition T = 0 on a source set represents the infimum travel time from the source to point \mathbf{x}, defined variationally as T(\mathbf{x}) = \inf_{\gamma} \int_{\gamma} n(\mathbf{r}) \, ds, where the infimum is taken over all admissible paths \gamma connecting the source to \mathbf{x}, and ds denotes the element along \gamma. This encodes the minimal "optical length" or time in with position-dependent speed $1/n(\mathbf{x}), generalizing classical measures to curved or anisotropic spaces. In uniform media, where the refractive index n(\mathbf{x}) is constant, the eikonal equation simplifies to |\nabla T| = n, yielding T(\mathbf{x}) = n \cdot d_E(\mathbf{x}), with d_E the standard from the source; this reflects the isotropic propagation at constant speed. For spatially varying n(\mathbf{x}), the formulation induces a Finsler or Riemannian on the domain, where the becomes ds = n(\mathbf{r}) \sqrt{d\mathbf{r} \cdot d\mathbf{r}} in the Riemannian case, defining geodesics as curves minimizing the integral and enabling shortest paths in heterogeneous environments. Such capture and inhomogeneity, with Finsler structures allowing direction-dependent lengths while preserving positive homogeneity. To handle discontinuities in n(\mathbf{x}) or non-smooth boundaries, the appropriate notion of solution is the viscosity solution, which ensures and for the eikonal equation as a Hamilton-Jacobi . Viscosity solutions guarantee that T(\mathbf{x}) coincides with the minimal arrival time, even across interfaces, by satisfying the equation in a weak sense via test functions at maxima and minima; this framework, developed by Crandall and Lions, resolves ambiguities in classical derivatives and confirms that the distance function is the unique continuous viscosity solution. This continuous shortest-path perspective links the eikonal equation to optimal theory, where the characteristics—or rays—emerge as optimal trajectories minimizing the time integral under velocity constraints. In the Hamilton-Jacobi-Bellman framework, the eikonal corresponds to a static case of the value function for a problem with unit-speed and cost n(\mathbf{x}), with lines tracing the minimizing paths.

Applications

Geophysics and Seismology

In geophysics and seismology, the eikonal equation is fundamental for modeling seismic wave propagation, particularly in travel-time tomography, where it enables the inversion of observed first-arrival times T to recover the subsurface velocity model v(\mathbf{x}) through the relation |\nabla T| = 1/v. This approach avoids explicit ray tracing by treating the traveltime field as a solution to a Hamilton-Jacobi equation, allowing efficient computation of wavepaths that broaden to account for finite-frequency effects and off-ray sensitivity. A key advancement is the wavepath eikonal traveltime inversion (WET) method, which unifies various tomographic schemes by back-projecting residuals along curved wavepaths derived from the eikonal solution, offering improved resolution over traditional straight-ray approximations while remaining computationally efficient—nearly an order of magnitude faster than full wave-equation methods. The eikonal equation also underpins ray-based techniques in seismic , where rays are traced as characteristics of the equation through heterogeneous, layered models to position reflections accurately. In prestack depth , solving the eikonal provides traveltimes for ray , enabling the construction of operators that handle complex velocity structures like salt domes or faults in sedimentary basins. This ray-theoretic framework, rooted in high-frequency asymptotics, facilitates the simulation of wave fronts and amplitudes, essential for constructing velocity models in exploration seismology. To address anisotropy prevalent in crustal rocks, such as shales or fractured reservoirs, the eikonal equation is extended to or elliptically media, where the depends on direction via parameters like Thomsen's \epsilon and \delta. In vertically (VTI), an acoustic sets the vertical shear-wave to zero, simplifying the eikonal to depend primarily on the normal-moveout and the \eta , which captures anellipticity and improves kinematic accuracy for P-wave with minimal . This formulation is crucial for and in anisotropic overburdens, enhancing image quality by correcting for stretching in tilted or elliptical wavefronts. For more general , factored forms of the eikonal allow numerical solutions that handle quasi-P and quasi-SV waves, vital for crustal inversion. In earthquake studies, the eikonal equation supports first-arrival picking by computing theoretical traveltimes for - and S-phases across networks, aiding automated detection and relocation in heterogeneous media. For instance, during the 2021 M_S 6.4 Yangbi earthquake in , eikonal-based inversion of multiple arrivals (, sPg, PmP) refined locations by incorporating velocity variations, reducing uncertainties compared to 1D layered models. Additionally, velocity inversion using eikonal-derived finite-frequency kernels—such as those from phase-front tracking of surface waves—maps crustal structure by weighting sensitivities to off-great-circle paths. These applications highlight the eikonal's role in bridging kinematic modeling and structural imaging for assessment.

Computer Vision and Graphics

In , the eikonal equation facilitates the fast computation of distance fields on and voxel grids, which are essential for analysis by yielding or weighted distances from arbitrary points to object boundaries. These fields, obtained by solving |\nabla d(\mathbf{x})| = 1/F(\mathbf{x}) where F denotes the local speed (often constant for distances), enable applications such as silhouette extraction and from point clouds, with errors as low as 3.89% compared to exact methods in benchmarks like the model. The Fast Marching Method (FMM), a seminal technique, achieves this in O(N \log N) time for N grid points by propagating a front, supporting weighted variants for anisotropic media in modeling and . For path planning and collision avoidance, the eikonal equation models optimal trajectories by treating obstacles as regions of zero speed (corresponding to infinite refractive index n = \infty), computing a travel-time field that guides agents around hazards via on the solution. In robotic , such as marine vessel routing, FMM-based solvers generate collision-free paths by expanding wavefronts from start points, incorporating dynamic obstacles through anisotropic speed functions that enforce distances (e.g., 200 m at the bow), with real-time replanning in complex environments like crowded ports. This approach avoids local minima by leveraging the eikonal's viscosity solution properties, ensuring globally optimal paths under spatial-temporal constraints. In computer graphics, the eikonal equation drives refraction simulation by deriving ray equations from |\nabla S| = n(\mathbf{x}), where n is the spatially varying refractive index, to trace light paths accurately in non-homogeneous media like glass or scattering fluids. The Eikonal Rendering technique, implemented on GPUs, approximates global illumination by adaptively marching wavefronts for caustics and internal reflections, achieving interactive frame rates (e.g., 24.8 fps for refractive wine glasses) while handling attenuation and anisotropy for realistic rendering of transparent objects. This method extends to broader light transport approximations, reducing computational overhead in scenes with complex refractive elements compared to full Monte Carlo ray tracing. The eikonal equation integrates seamlessly with level-set methods for evolving interfaces in , where solutions provide signed distance functions to reinitialize the level-set \phi and propagate fronts at speeds modulated by image gradients (e.g., F = 1 - |\nabla I|). In applications like , competing eikonal-derived distance fields from interior and exterior seeds delineate regions such as nodes, enabling topology changes and robust detection without explicit parameterization, often via fast marching for O(N \log N) efficiency. This front propagation framework, as in the Chan-Vese model, minimizes energy functionals for accurate segmentation of noisy data like MRI tumors.

Numerical Solution Techniques

Grid-Based Discretization Methods

Grid-based methods approximate solutions to the eikonal equation on structured Cartesian grids by replacing the continuous with approximations, typically ensuring and through upwind schemes that respect the directional of wavefronts. These methods discretize the into a uniform grid with spacing h in each direction, where the arrival time T is computed at grid points (x_i, y_j) = (ih, jh), and the slowness n(x) is evaluated at the same points. The core challenge is to approximate |\nabla T| in a way that maintains the monotonicity of the solution, avoiding oscillations and ensuring to the physically relevant . In two dimensions, a basic upwind approximation uses backward differences to estimate the components, assuming the advances from known values with smaller T. For instance, at grid point (i,j), the scheme solves \sqrt{ \left( \frac{T_{i,j} - T_{i-1,j}}{h} \right)^2 + \left( \frac{T_{i,j} - T_{i,j-1}}{h} \right)^2 } = n_{i,j} for T_{i,j}, which corresponds to an upwind in the first quadrant and enforces by depending only on previously computed upwind values. This formulation arises from the continuous eikonal equation |\nabla T| = n(x), where the structure preserves the of the . More general upwind schemes consider all four quadrants by taking the minimum over possible orientations to handle arbitrary directions. Godunov's scheme enhances this by providing a monotone approximation through a Godunov-type flux that selects upwind differences based on the local gradient signs, ensuring the numerical Hamiltonian is non-decreasing and the solution remains causal with respect to wavefront propagation. Specifically, the scheme approximates the gradient as |\nabla T_{i,j}| \approx \sqrt{ \left( \max\left( D_x^- T_{i,j}, 0 \right) - \min\left( D_x^+ T_{i,j}, 0 \right) \right)^2 + \left( \max\left( D_y^- T_{i,j}, 0 \right) - \min\left( D_y^+ T_{i,j}, 0 \right) \right)^2 }, where D_x^\pm and D_y^\pm are the backward and forward differences in the x and y directions, respectively, set equal to n_{i,j}. This construction guarantees the scheme is monotone, as the update for T_{i,j} takes the minimum of the current value and the solved value from the scheme, ensuring non-increasing arrival times and preventing non-physical values. Extensions to n dimensions generalize the upwind approximation by including squared differences along all coordinate axes, yielding \sqrt{ \sum_{k=1}^n \left( \max\left( D_k^- T, 0 \right) - \min\left( D_k^+ T, 0 \right) \right)^2 } = n, which maintains consistency for the full tensor product grid. For computational efficiency in high dimensions, dimensional splitting decomposes the operator into sequential one-dimensional solves along each axis, reducing the nonlinearity while preserving first-order accuracy. Error analysis for these schemes demonstrates first-order accuracy, with the local truncation error bounded by O(h) due to the one-sided differences, leading to global convergence at rate O(h) under uniform grid refinement. Monotonicity and consistency ensure that, as h \to 0, the numerical solution converges uniformly to the unique viscosity solution of the eikonal equation, as established by the Crandall-Lions theory for Hamilton-Jacobi equations.

Upwind and Monotone Schemes

Upwind and monotone schemes represent a class of efficient numerical methods for solving the eikonal equation, particularly suited for large-scale problems where causality and solution monotonicity must be preserved to ensure accurate front propagation without spurious oscillations. These schemes build on finite difference discretizations by enforcing an upwind ordering that respects the direction of information flow, akin to characteristics in hyperbolic equations, thereby achieving high speed and stability for the static Hamilton-Jacobi formulation of the eikonal. The fast marching method, introduced by Sethian in , is a seminal Dijkstra-like algorithm that solves the eikonal equation by sorting points according to tentative arrival times T and updating neighboring points using the solved eikonal values in a single forward pass. This approach treats the discrete domain as a where each point is a , and edges connect to upwind neighbors; a efficiently selects the next point with the smallest tentative T, freezing it once accepted to maintain causality. The method's efficiency stems from its O(N log N) in d dimensions, where N is the number of points, making it particularly effective for domains by avoiding iterative relaxations. Fast sweeping methods, developed by Zhao in 2005, provide an alternative iterative approach using the same upwind discretizations, such as Godunov schemes, combined with Gauss-Seidel relaxations and multiple directional sweeps across the grid to propagate information along characteristics. Unlike fast marching, fast sweeping does not require sorting and achieves practical O(N) complexity with a fixed number of sweeps (typically 4 in 2D, more in higher dimensions), making it simpler to parallelize and suitable for inhomogeneous media. Convergence is ensured by the monotonicity of the updates, where each grid point's value is non-increasing until the viscosity solution is reached. Ordered upwind methods extend the fast marching paradigm to more general static Hamilton-Jacobi equations, including anisotropic eikonal variants, by using structured priority queues to perform multi-dimensional updates in a causal order. Developed by Sethian and Vladimirsky in 2000, these methods divide the grid into ordered simplices and solve local systems along characteristic directions, ensuring single-pass computation with O(N log N) complexity even for non-isotropic speeds. This allows for accurate handling of complex geometries and varying indices without repeated sweeps. A key feature of both fast and ordered upwind schemes is their monotonicity preservation, which guarantees to the unique viscosity of the eikonal equation as a static Hamilton-Jacobi equation by ensuring that the numerical is non-decreasing in the values. This prevents oscillations and ensures stability, as upwind differencing selects only known (upwind) information, aligning with the for the eikonal. Such monotonicity is crucial for applications requiring precise travel-time fields, like seismic modeling. In terms of runtime comparisons for domains, fast marching methods typically outperform direct sparse linear solvers, which discretize the eikonal into a solvable via methods like multigrid (O(N)) or direct factorization (O(N^{1.5})), especially for large N exceeding 10^6 points; this efficiency gap widens in heterogeneous media, where iterative solvers may require hundreds of relaxations, while fast marching achieves optimal single-pass performance.

References

  1. [1]
    [PDF] Math 5587 – Lecture 21
    Nov 16, 2016 · This nonlinear partial differential equation is called the eikonal equation. The eikonal equation also shows up in wave propagation ...
  2. [2]
    [PDF] GEOMETRICAL OPTICS
    Equation (1.1.5) is often called the eikonal equation or the ray-tracing equation. Many of the waves that we have already study satisfy this equation, such as ...
  3. [3]
    The eikonal equation - Book chapter - IOPscience - Institute of Physics
    The eikonal equation is a partial differential equation with non-linearity found in wave propagation. It is an approximated version of the wave equation.
  4. [4]
    [PDF] Chapter 2 Geometrical optics - MIT OpenCourseWare
    We now give a general solution of the eikonal equation, albeit in a somewhat implicit form, in terms of rays. The rays are the characteristic curves for the.
  5. [5]
    [PDF] Mathematical Morphology and Eikonal Equations on Graphs ... - HAL
    Mar 3, 2009 · Our formulation of the eikonal leads to novel applications of this equation such as nonlocal image segmentation and weighted distances or data ...<|control11|><|separator|>
  6. [6]
    [PDF] A level set-based Eulerian approach for anisotropic wave propagation
    For isotropic wave propagations, the eikonal equations, which are of Hamilton–Jacobi form, are usually solved using viscosity solution-based solvers over ...
  7. [7]
    [PDF] Wave propagation and ray methods - Arizona Math
    WKB approximation u ∼ exp(i(ωt − θ(x))/)[u0(x, y) + u1(x, y) + ...] leads ... The eikonal equation. The equation |∇θ|2 = µ2(x) is nonlinear and ...
  8. [8]
    Principles of Optics. Second (Revised) Edition. : Born, Max, & Emil ...
    Oct 17, 2023 · Principles of Optics. Second (Revised) Edition. by: Born, Max, & Emil Wolf,. Publication date: 1964-01-01. Publisher: Pergamon Press. Collection ...Missing: official | Show results with:official
  9. [9]
    [PDF] Lecture 10 Approximation methods: WKB method - TCM
    In optics, WKB is known as eikonal method, and in general referred to by mathematicians as short wavelength asymptotics. In quantum mechanics, it provides ...
  10. [10]
    [PDF] The WKB approximation and Ray Theory - Biello
    We will use the Eikonal equation to construct the ODEs for the rays and we will use the Amplitude equation to describe how the amplitude of the waves changes as ...
  11. [11]
    3.Fermat's Principle of Least Time - Galileo and Einstein
    Fermat famously stated in the 1630's that a ray of light going from point A to point B always takes the route of least time.
  12. [12]
    The principle of least time - SEG Wiki
    Apr 26, 2021 · Pierre de Fermat (1601–1665) formulated the rule known as Fermat's principle of least time. In his original statement, Fermat asserted that ...
  13. [13]
    Descartes on the Refraction and the Velocity of Light
    Five months later, Clerselier dispatched two letters to Fermat in which the latter's assumption-the principle of least time-was attacked and torn to shreds. The ...<|separator|>
  14. [14]
    Detailed derivation of the generalized Snell–Descartes laws from ...
    Mar 6, 2023 · Beginning with Fermat's principle, we provide a detailed derivation of the generalized laws of refraction and reflection for a geometry ...
  15. [15]
    [PDF] The Eikonal - Neo-classical physics
    HEINRICH BRUNS. REGULAR MEMBER OF THE SAXON SOCIETY OF SCIENCE ... will use the term “eikonal” in order to have a concise expression. Each map ...
  16. [16]
    Applications of the eikonal approximation in quantum mechanical ...
    May 1, 2023 · The term “eikonal” was coined by Heinrich Bruns in his 1895 manuscript “Das Eikonal” published in Leipzig by S. Hirzel, an English ...
  17. [17]
    (PDF) Eikonal Functions: Old and New - ResearchGate
    Jan 1, 2025 · ... Bruns, unaware of Hamilton's work, proposed a similar. idea [Bruns, 1895]. Bruns also coined the name 'eikonal' to the same. function called ...
  18. [18]
    [PDF] On some Results of the View of a Characteristic Function in Optics ...
    Hamilton divides mathematical optics into two principal parts: one part proposing to find in every particular case the form of the characteristic function V , ...Missing: eikonal | Show results with:eikonal
  19. [19]
    Hamilton's Method in Geometrical Optics - Optica Publishing Group
    Hamilton's method. In 18325 Hamilton defined the characteristic function V of an optical system as the optical length of the ray joining ...Missing: primary source
  20. [20]
    Hamilton-Jacobi equation - Scholarpedia
    Oct 21, 2011 · The Hamilton-Jacobi equation is used to generate particular canonical transformations that simplify the equations of motion.
  21. [21]
    [PDF] On the Geometry of the Hamilton-Jacobi Equation - ICMAT
    With an emphasis on mechanics, Carl Gustav Jacob Jacobi (1804 − 1851) deepened. Hamilton's formulation, by clarifying some mathematical issues, and ...<|separator|>
  22. [22]
    Einstein and Hilbert: The Creation of General Relativity - arXiv
    Apr 25, 2005 · Recent controversy, raised by a much publicized 1997 reading of Hilbert's proof-sheets of his article of November 1915, is also discussed.
  23. [23]
    [PDF] Einstein and Hilbert: Two Months in the History of General Relativity
    In his postcard of November 7, 1915, EINSTEIN tells HILBERT that he is ... HILBERT-EINSTEIN equations, it seems to us proper to call the general theory.Missing: eikonal | Show results with:eikonal<|separator|>
  24. [24]
    Geometrical theory of optical imaging (IV) - Principles of Optics
    Since S(r) satisfies the eikonal equation §3.1 (15), this function is fully specified by the refractive index function (r) alone, together with the appropriate ...
  25. [25]
    Eikonal Functions: Old and New - SpringerLink
    Eikonal functions are among the oldest and most useful tools in optics. They form the foundations of geometrical optics and optical aberrations theory.
  26. [26]
    None
    ### Summary of Sections from Notes on Geometrical Optics
  27. [27]
    [PDF] Introduction to Partial Differential Equations, Math 463/513, Spring ...
    Apr 10, 2015 · 1.7.1 Derivation of the Eikonal Equation . ... Equation (13.10) is called Snell's law. 13.3 Electromagnetic Waves in the Atmosphere. The ...
  28. [28]
    [PDF] The Wave Equation - MIT OpenCourseWare
    Mar 3, 2007 · Fermat's Principle implies Snell's Law. The travel time curve, plotted as a function of offset, is typically a hyperbolic function. Near ...Missing: derivation | Show results with:derivation
  29. [29]
    [PDF] 1207.1.K.pdf - Caltech PMA
    This equation is known in optics as the eikonal equation. It is formally the same as the. Hamilton-Jacobi equation of classical mechanics3 if we identify Ω ...
  30. [30]
    [PDF] Geodesic distances
    The eikonal equation is simply a reflection of the fact that distance functions grow with constant speed radially away from the center point. It is a ...
  31. [31]
    [PDF] fast sweeping methods for static Hamilton-Jacobi equations
    A special case of this type of equations is the eikonal equation: (3) ... An important application for (2) is obtaining geodesic distance on a manifold.
  32. [32]
    [PDF] Efficient Fast Marching with Finsler metrics. - HAL
    Sep 28, 2012 · The Riemannian metric is equal to the euclidean norm, except on a band of width 1/100 along a spiraling curve Γ where it has eigenvalues 1 ...Missing: media | Show results with:media
  33. [33]
    viscosity solutions of hamilton-jacobi equations1
    VISCOSITY SOLUTIONS OF HAMILTON-JACOBI EQUATIONS1. BY. MICHAEL G. CRANDALL AND PIERRE-LOUIS LIONS. Abstract. Problems involving Hamilton-Jacobi equations—which ...
  34. [34]
    [PDF] viscosity solutions of the eikonal equations - UChicago Math
    The eikonal equation is a nonlinear PDE related to wave propagation. This paper will investigate how distance functions are viscosity solutions of eikonal ...
  35. [35]
    [PDF] An overview of static Hamilton-Jacobi equations - UCSB Math
    This paper reviews some of the existence and uniqueness techniques for the time independent cases with a partic- ular interest in Eikonal-like Equations. In ...
  36. [36]
  37. [37]
    [PDF] 3D Distance Fields: A Survey of Techniques and Applications
    The Eikonal equation states the obvious inverse relationship between the speed of the front and the gradient of the arrival time. Since F does not have to be ...
  38. [38]
    [PDF] A Schrödinger wave equation approach to the eikonal ... - UF CISE
    The eikonal equation has several applications in image analysis, viz. signed distance functions for shape silhouettes, surface reconstruction from point clouds ...<|separator|>
  39. [39]
    Marine Applications of the Fast Marching Method - Frontiers
    Jan 27, 2020 · In this paper, several solutions based on the Fast Marching Method are proposed. The basic method focus on collision avoidance and optimal planning.
  40. [40]
    A Path Planning Method for Ship Collision Avoidance Considering ...
    This research proposes a novel path planning method based on the fast marching method to specifically assist with safe navigation for autonomous ships.A Path Planning Method For... · 3. A Proposed Approach · 3.3. 1. Trajectory...
  41. [41]
    [PDF] Eikonal Rendering: Efficient Light Transport in Refractive Objects
    To summarize, we present a new fast and versatile framework de- rived from the eikonal equation that can jointly reproduce many lighting effects around complex ...
  42. [42]
    [PDF] Efficient Segmentation Based on Eikonal and Diffusion Equations
    Segmentation uses Eikonal and diffusion equations to solve PDEs, computing distance functions for each region. A competition between these functions determines ...
  43. [43]
    [PDF] level set methods and their applications in image science
    Level set methods are used in image science, especially for segmentation, processing level lines, and representing boundaries, often with PDE techniques.
  44. [44]
    A second-order distributed memory parallel fast sweeping method ...
    Feb 1, 2023 · This is called a Godunov upwind difference scheme [4]. Using this discretization scheme in the Eikonal equation results in a large nonlinear ...
  45. [45]
    [PDF] Lecture notes on viscosity solutions
    These notes are concerned with viscosity solutions for fully nonlinear equa- tions. A majority of the notes are concerned with Hamilton-Jacobi equations.
  46. [46]
  47. [47]
    A fast marching level set method for monotonically advancing fronts.
    A fast marching level set method is presented for monotonically advancing fronts, which leads to an extremely fast scheme for solving the Eikonal equation.
  48. [48]
    Ordered upwind methods for static Hamilton–Jacobi equations | PNAS
    We introduce a family of fast ordered upwind methods for approximating solutions to a wide class of static Hamilton–Jacobi equations with Dirichlet boundary ...
  49. [49]
    [PDF] arXiv:1403.1937v2 [math.NA] 8 Feb 2015
    Feb 8, 2015 · Our approach is straightforward to implement (by deploying any sparse linear solver) and holds its own against contemporary fast marching ...