Fact-checked by Grok 2 weeks ago

Optical path

The optical path, also known as the optical path length (OPL), is a fundamental quantity in optics that describes the effective propagation distance of light through a medium, defined as the line integral of the refractive index n along the geometric path ds traversed by a light ray: \mathrm{OPL} = \int n \, ds. This measure is physically equivalent to the distance light would travel in vacuum during the same time, given by c \times t, where c is the speed of light in vacuum and t is the travel time. In a homogeneous medium with constant refractive index, the OPL simplifies to the product of n and the physical path length L, i.e., \mathrm{OPL} = nL. The concept of optical path underpins Fermat's principle, which states that light rays propagate along paths where the OPL is stationary—typically a minimum or maximum with respect to small variations in the trajectory—ensuring the principle of least time for light travel between two points. This principle, originally formulated by in 1662, derives key laws of geometric optics, such as reflection and refraction (), by minimizing the OPL for ray paths at interfaces between media. In inhomogeneous media, where n varies spatially (e.g., in graded-index materials or atmospheric ), the OPL accounts for bending or distortion of rays, making it essential for analyzing phenomena like mirages or aero-optical effects in high-speed flows. Beyond geometric optics, the optical path plays a critical role in wave optics and , where differences in OPL determine phase shifts and interference patterns; for instance, in a , inserting a sample in one arm alters the OPL, shifting fringes whose count quantifies the change with high precision. Applications extend to optical imaging systems, where equalizing OPLs across rays ensures aberration-free focus, as in lens design and ray tracing algorithms. In for telescopes, real-time adjustments compensate for atmospheric OPL variations to sharpen stellar images. Additionally, OPL measurements enable techniques like low-coherence for in random media, revealing path-length distributions in environments such as biological tissues. In interferometric arrays like the telescope, OPL equalizers maintain phase coherence across baselines for high-resolution astronomical imaging.

Fundamentals

Definition

The optical path represents the effective distance travels through a medium, equivalent to the distance it would traverse in to accumulate the same . This measure accounts for the medium's influence on 's propagation speed, providing a way to quantify the cumulative effect on the wavefront's advance as if were unimpeded in free space. The originated in 17th-century , rooted in Pierre de Fermat's 1657 principle of least time, which posits that follows the path minimizing travel time and, by extension, the . provided early recognition of this idea within wave theory in his 1678 Traité de la Lumière, where he modeled propagation as secondary wavelets advancing at speeds inversely proportional to the , implicitly incorporating optical path considerations for construction. In contrast to the geometric path, which denotes the purely physical along the ray's trajectory regardless of the medium, the optical integrates the to reflect the actual slowing of , yielding a longer effective distance in denser materials. For instance, in where the n = 1, the optical coincides exactly with the geometric , while in air with n \approx 1, the two are virtually indistinguishable for typical distances.

Optical Path Length

The optical path length (OPL), often denoted as \Lambda, quantifies the effective distance traveled by light through a medium by accounting for the medium's refractive index, providing a vacuum-equivalent measure that determines the accumulated phase of the light wave. It is defined mathematically as the line integral along the ray path: \Lambda = \int n \, ds, where n is the refractive index at each point along the infinitesimal path element ds. This formulation extends the geometric path length to incorporate the slowing of light in denser media, making OPL a fundamental quantity in wave and ray optics. Physically, the OPL corresponds to the time \tau light takes to traverse the path, since the speed of light in the medium is c/n (with c the vacuum speed), yielding \tau = \int (n \, ds)/c = \Lambda / c. This equivalence arises because the phase advance \phi = (2\pi / \lambda) \Lambda (where \lambda is the vacuum wavelength) directly governs interference and diffraction phenomena, emphasizing OPL's role in phase accumulation without altering the underlying propagation time. In applications like , OPL ensures that phase shifts are predicted accurately even in varying environments. The units of OPL are meters, identical to the geometric path length, facilitating direct comparison with vacuum propagation distances. A key property in ray optics is that OPL remains invariant under coordinate transformations along the ray trajectory, preserving its scalar nature and utility in geometric optics formulations. This invariance underpins its connection to , where stationary OPL paths correspond to actual light rays.

Mathematical Formulation

Homogeneous Media

In homogeneous media, where the n remains constant along the path of propagation, the (OPL) simplifies to a straightforward product of the refractive index and the geometric path length L. This is expressed mathematically as \text{OPL} = n \times L, where L is the physical distance traveled by the ray in the medium. This formulation assumes a uniform medium without spatial variations in n, allowing to propagate in lines according to the principles of . The concept of optical path length originates from the time-of-flight perspective in . In such a medium, the speed of light is v = c / n, where c is the in . Thus, the time t for to traverse the L is t = \frac{L}{v} = \frac{L}{c / n} = \frac{n L}{c}. This travel time is equivalent to the time it would take for to cover a distance n L in , establishing the as an effective vacuum-equivalent path that accounts for the medium's slowing effect on . This equivalence underpins for path minimization in uniform media. This simplified expression applies under key assumptions: the medium must be isotropic (with n independent of light direction), non-absorbing (no energy loss that alters the effective path), and free of refractive index gradients that would curve the light ray. These conditions hold for many common materials like air (n \approx 1) or crown glass in basic optical setups. For instance, consider a ray passing through a slab of glass with thickness L = 1 cm and refractive index n = 1.5. The OPL is then $1.5 \times 1 cm = 1.5 cm, representing the effective path length as if the light had traveled 1.5 cm in vacuum. This calculation is fundamental in designing simple lenses or windows where uniformity is assured.

Inhomogeneous Media

In inhomogeneous media, where the refractive index n varies spatially, the (OPL) is computed by integrating along the actual ray path, contrasting with the simple product used in uniform media. This variation in n(\mathbf{r}) arises from gradients due to factors like or , requiring a more complex evaluation to capture the true propagation delay. The general formula for the between points A and B is given by the \text{OPL} = \int_A^B n(\mathbf{r}) \, ds, where ds is the differential arc length along the ray path. This integral accounts for the local refractive index at each point, providing the effective distance light travels as if in vacuum. For practical computation in media with complex index gradients, such as graded-index (GRIN) fibers, numerical methods like ray tracing software approximate the integral by discretizing the path into segments and summing contributions. These tools solve differential equations governing ray propagation, enabling simulations of light guiding in multimode fibers where the index decreases parabolically from the core center. Within the eikonal approximation, valid for high-frequency light where wavefronts are nearly planar locally, light rays follow curved paths that minimize the OPL according to . The |\nabla S| = n(\mathbf{r}), where S is the optical path function, governs this bending, ensuring rays take stationary paths through varying media. This formulation is essential in , where temperature-induced index gradients cause mirages by bending rays toward denser air layers, creating illusory images of distant objects.

Optical Path Difference

The optical path difference (OPD), denoted as Δ, is defined as the difference between the optical path lengths (OPLs) of two light rays or beams traveling along different routes from a common source to a common observation point. This quantity accounts for both the physical distances traversed and the refractive indices of the media involved, providing a measure of the effective path disparity that influences wave superposition. For two paths in potentially different media, the OPD is given by the formula \Delta = n_1 L_1 - n_2 L_2, where n_1 and n_2 are the refractive indices, and L_1 and L_2 are the physical path lengths along each route, respectively. In vacuum or air (where n \approx 1), this simplifies to the geometric path difference. The OPD directly determines the phase difference \delta between the two waves, expressed as \delta = \frac{2\pi}{\lambda} \Delta, where \lambda is the wavelength of the light in vacuum. This phase shift governs interference phenomena: constructive interference occurs when \Delta = m \lambda for integer m, resulting in maximum intensity as the waves reinforce each other; destructive interference arises when \Delta = (m + \frac{1}{2}) \lambda, leading to minimum intensity due to wave cancellation. These conditions assume coherent, monochromatic sources and equal amplitudes for the beams. A prominent example of OPD's role is in the , where a is split into two perpendicular paths by a partially reflecting mirror and reflected back by movable mirrors before recombination. The OPD here is \Delta = 2(d_1 - d_2), accounting for the round-trip travel in each arm (assuming air, n=1), with d_1 and d_2 as the distances to the mirrors. Displacing one mirror by a distance x changes the OPD by $2x, producing interference fringes that shift across the observation plane; a displacement of \lambda/2 corresponds to a full fringe shift, enabling precise measurements of length or variations.

Fermat's Principle

Fermat's principle asserts that a ray of light traveling between two fixed points follows the path for which the optical path length is stationary, corresponding to an extremum—typically a minimum—with respect to small variations in the path. This variational principle underpins ray optics by selecting the trajectory that light actually takes among all possible paths. The principle was first proposed by the French mathematician in a 1662 letter to Cureau de la Chambre, as a means to explain the of . Although formulated prior to the advent of wave optics in the late 17th and 19th centuries, it remains fully consistent with wave-based descriptions of light propagation, such as Huygens' principle. Mathematically, is expressed through the , requiring that the —the of the n along the path—be stationary: \delta \int n \, ds = 0, where ds is the differential along the path and the is taken between the two points. This condition, applied to paths involving interfaces between media, yields the law of reflection (equal angles of incidence and reflection) and of (n_1 \sin \theta_1 = n_2 \sin \theta_2) as consequences of minimization. The referenced here is elaborated in the section on inhomogeneous media.

Factors Affecting Optical Path

Refractive Index

The n of a medium is defined as the ratio of the in c to its speed in the medium v, expressed as n = \frac{c}{v}. This quantifies how much slower propagates through the material compared to . For most materials, n > 1, reflecting the reduced velocity due to the medium's properties. Physically, the originates from interactions between the 's and the electrons in the medium's atoms or molecules. The oscillating field induces , creating oscillating dipoles that radiate secondary waves; the of these with the incident effectively delays propagation. Higher material density increases the by providing more electrons per unit volume for these interactions. Variations in environmental conditions also affect the . The temperature coefficient \frac{dn}{dT} measures its sensitivity to temperature changes, which can induce thermal lensing in optical systems where nonuniform heating creates refractive index gradients acting like a . Pressure influences arise similarly through changes; for instance, standard atmospheric air has a refractive index of approximately 1.0003 at 15°C and 101.325 kPa. In media that absorb light, the refractive index becomes complex, n = n_r + i k, where the real part n_r governs the phase shift and thus the optical path in non-absorptive scenarios, while the imaginary part k describes . This refractive index fundamentally scales the optical path length in homogeneous media.

Wavelength Dependence

The (OPL) of propagating through a medium is inherently dependent on the due to material dispersion, where the n varies as a function of \lambda, denoted as n(\lambda). This dependence arises from the interaction of with the material's structure, leading to different velocities for various colors of and complicating the design of optical systems that must handle or polychromatic . In normal dispersion, which characterizes most transparent media like in the , the increases with decreasing , meaning shorter wavelengths (e.g., ) experience a higher n than longer wavelengths (e.g., red light). As a result, for a fixed physical length through such a dispersive medium, the OPL is effectively longer for than for red light, since OPL is the product of the and the geometric . This differential OPL contributes to phenomena like chromatic in optical fibers and lenses, where signals spread temporally or spatially. The provides an empirical model for describing n(\lambda) in dispersive materials, derived from a classical oscillator of the bound electrons responding to the of light. Developed by Wolfgang Sellmeier in 1871, it fits experimental data by summing contributions from multiple terms associated with the material's and oscillators, offering accurate predictions of across a wide spectral range without invoking . However, near regions of strong —such as electronic transition bands in the or vibrational bands in the —materials exhibit anomalous , where dn/d\lambda > 0, causing the to decrease with decreasing . This counterintuitive behavior, first noted in the , stems from the rapid variation in the real part of the complex adjacent to lines, and it can lead to unusual effects in or resonant optical applications.

Applications

Interferometry

In interferometry, the optical path difference (OPD) plays a central role in generating fringes by introducing controlled shifts between traveling along distinct paths. These fringes arise when the OPD between interfering beams results in constructive or destructive , enabling high-precision measurements of displacements, refractive indices, and other . In the Fabry-Pérot interferometer, consisting of two parallel partially reflecting mirrors forming an , multiple reflections create successive beams with incremental OPDs of $2nd\cos\theta, where n is the , d is the mirror separation, and \theta is the incidence angle. Constructive occurs when this OPD equals an multiple of the , producing peaks or minima that form sharp fringes, allowing of lines down to picometer scales. Similarly, the Mach-Zehnder interferometer splits a into two arms with adjustable path lengths, recombining them to produce fringes sensitive to OPD variations as small as a fraction of a ; the resulting shift manifests as fringe shifts or modulations, making it ideal for dynamic sensing applications like vibration analysis. A key application of OPD in is wavelength measurement, where the fringe spacing or provides direct . For the m-th , the condition for maximum is given by \Delta = m\lambda, allowing \lambda to be determined from the measured OPD \Delta (via arm displacement or tuning) and observed fringe count m, with accuracies exceeding 1 part in $10^8 in stabilized setups. Historically, the Michelson-Morley experiment of 1887 employed an interferometer to detect the luminiferous ether by measuring expected shifts from Earth's motion through it, which would alter the OPD in arms due to velocity-dependent light propagation. No such shift was observed, nullifying the ether hypothesis and supporting relativistic principles, with the apparatus achieving OPD sensitivities of about 0.01 fringes. In modern holography, OPD governs the recording and reconstruction of three-dimensional images by capturing the wavefront phase variations from an object, encoded in the interference pattern between object and reference beams. During reconstruction, the hologram diffracts light to replicate these OPDs, restoring the original wavefront and enabling parallax-based 3D visualization without computational aids in classical setups.

Optical Design

In optical design, the optimization of optical path length (OPL) is fundamental to achieving aberration-free imaging in lens systems, telescopes, and microscopes. Ray tracing techniques simulate the propagation of light rays through optical elements, calculating the OPL as the integral of the refractive index along each ray's path to evaluate and minimize aberrations. For instance, spherical aberration is corrected by ensuring that the OPL for paraxial and marginal rays from an object point to the image plane is equal, preventing wavefront distortion. Similarly, chromatic aberration is addressed by balancing OPL variations across wavelengths using materials with appropriate dispersion properties. These simulations enable designers to iteratively refine surface curvatures, thicknesses, and separations for high-performance systems. A key principle in optical design is the conservation of , which quantifies the throughput of an optical system as n^2 times the product of the beam's cross-sectional area and the it subtends (for homogeneous media). Along the optical path in lossless, reversible systems, etendue remains invariant, imposing fundamental limits on light collection and concentration. This conservation guides the design of and illumination , ensuring that parameters do not degrade due to mismatches in area or angle, as seen in objectives where maximizing etendue enhances light-gathering efficiency without introducing path-induced losses. Telecentric designs exemplify OPL optimization by positioning the aperture stop at the focal plane, making chief rays parallel to the and thus ensuring equal OPL from the object to the across the field of view. This equality minimizes and errors, ideal for precision metrology in microscopes. In fiber optics, OPL mismatches among propagating modes in multimode fibers cause , where rays following longer paths arrive later, broadening pulses and limiting data rates to below 1 Gbit/s over kilometer distances.

Materials and Examples

Common Materials

In optical path calculations, the n of a is a fundamental parameter that determines the effective path length through the medium. Common materials in are selected for their transparency and well-characterized , with refractive indices typically measured at the sodium D-line of 589 nm unless otherwise specified. These values serve as baselines for designing optical systems where the (OPL) is given by n \times d, with d as the physical thickness. Air and are standard reference media, with n \approx 1.000 for and n_{\text{air}} \approx 1.000277 at (), making the difference negligible for most laboratory and engineering applications (approximately $2.7 \times 10^{-4}). This small deviation arises primarily from air's density and composition but does not significantly affect in typical setups. are ubiquitous in optical elements like lenses and prisms. glass, often made from soda-lime compositions, has a refractive index of about 1.52 at 589 nm and a high (typically 50–60), indicating low suitable for achromatic designs. , with higher lead content, exhibits n \approx 1.62 at the same wavelength and a lower (around 30–40), which introduces greater chromatic but is valuable for correcting aberrations in compound lenses. Other everyday materials include , with n = 1.33 at 589 nm, commonly encountered in aqueous optics or biological . Diamond, prized for its hardness and clarity, has a high n = 2.42 at the same wavelength, leading to significant in gem applications. Polymers such as (polymethyl methacrylate) offer n \approx 1.49 and are favored for lightweight, cost-effective components in displays and protective covers.
MaterialRefractive Index (n at 589 nm)Abbe Number (if applicable)Typical Use
Vacuum1.000N/AReference medium
Air (STP)1.000277N/A
1.33N/ALiquid lenses,
(PMMA)1.49~57Lenses, windows
Crown Glass1.5250–60Achromatic doublets
1.6230–40Aberration correction
2.42~55High-index prisms, tools
These indices highlight how material choice influences OPL, with higher n values extending the effective path and enabling compact designs, as detailed in the refractive index section.

Calculation Examples

Consider a straightforward calculation of the optical path length (OPL) through a homogeneous medium, such as a crown glass slab with refractive index n = 1.5 and physical thickness d = 10 cm. The OPL, denoted as \Lambda, is computed as \Lambda = n d = 1.5 \times 10 cm = 15 cm, effectively equivalent to the distance light would travel in to match the phase delay experienced in the glass. In the context of Young's double-slit experiment conducted in air, where the refractive index is approximately 1, an optical path difference (OPD) of \Delta = 2 μm between the light paths from the two slits results in a phase shift that shifts the interference pattern. This OPD corresponds to constructive or destructive interference depending on the wavelength \lambda, with adjacent fringes separated by a path difference change of \lambda, or \lambda/2 for transitions between bright and dark fringes. For media with a graded refractive index, such as a linear profile where n varies from n_1 to n_2 over path length L, the OPL \Lambda = \int_0^L n(s) \, ds requires numerical approximation. The simple trapezoidal rule divides the path into N segments of width \Delta s = L/N, yielding \Lambda \approx \sum_{i=1}^N \frac{n_i + n_{i+1}}{2} \Delta s, which provides a reliable estimate for smooth gradients by treating each segment as a trapezoid. In homogeneous cases, error analysis reveals that uncertainties propagate directly: a 1% relative uncertainty in the n induces a 1% relative error in the , assuming the physical path length is precisely known, as derived from standard uncertainty propagation for the product \Lambda = n d.

References

  1. [1]
    [PDF] MITOCW | Lec 3 | MIT 2.71 Optics, Spring 2009
    Optical path length equals the integral along the ray trajectory of the index of refraction times dl. So what this means now in our case is that the optical ...<|control11|><|separator|>
  2. [2]
    [PDF] Optical Sensor Technology - University of Washington
    The optical path length of a ray traveling from point A to point B is defined as c times the time it takes the ray to travel from A to B. Assume a ray ...
  3. [3]
    General Terms - Exploring the Science of Light
    Optical fiber. Optical fiber. Optical path length – The product of the geometrical distance and the refractive index, in a medium of constant index of ...
  4. [4]
    [PDF] Fermat's Principle and the Laws of Reflection and Refraction ( )2
    We calculate the length of each path and divide the length by the speed of light to determine the time required for the light to travel between the two points.
  5. [5]
    [PDF] Winter 2008 When light travels from one medium to another, it
    And so the proper formulation of Fermat's principle is that light travels along paths of extremal optical path length. Typically these are minimal OPL paths, ...
  6. [6]
    [PDF] Physics and Computation of Aero-Optics - SciSpace
    as the optical path length (OPL). It is most commonly derived from geometric optics by assuming straight ray paths and is generally dependent on the flow ...
  7. [7]
    [PDF] Lecture 41 - Purdue Physics department
    Inserting something in one path changes the optical path length and shifts the interference fringes. – Counting the number of fringes that shift determines the ...
  8. [8]
    [PDF] Optical Imaging
    This approximation emphasizes the path travelled by light through media to find the locations and sizes of images.
  9. [9]
    [PDF] REAL TIME PHASE CORRECTION OF OPTICAL IMAGING SYSTEMS
    The phase or optical path length in each section of the aperture is changed sequentially and the effect on the energy density of the image or image sharpness ...
  10. [10]
    Optical path-length spectroscopy of wave propagation in random ...
    Apr 1, 1999 · By using a backscattering technique based on low-coherence interferometry, we are able to determine the optical path-length distribution for ...
  11. [11]
    [PDF] Optical Path Length Equalization - CHARA Array
    The Optical Path Length Equalizer (OPLE) subsystem provides the variable delays nec- essary to keep the telescopes phased as they track an object under ...
  12. [12]
    Optical Path Length – optical phase, Fermat's principle - RP Photonics
    It refers to the product of the physical path length that light travels through a medium and the refractive index of that medium.
  13. [13]
    History of Geometric Optics
    The earliest surviving optical treatise, Euclid's Catoptrics 1 (280BC), recognized that light travels in straight-lines in homogeneous media.
  14. [14]
    [PDF] Huygens principle; young interferometer; Fresnel diffraction
    – the “history” of where the field has been through. • distance traveled. • materials in the path. • generally, the “optical path length” is inscribed in the ...
  15. [15]
    [PDF] Geometrical Optics - Physics Internal Website
    When a light ray travers- es an optical system consisting of several homogeneous media in sequence, the optical path is a sequence of straight-line segments.
  16. [16]
    [PDF] Introduction to Optoelectronic Engineering,
    time to travel distance d in a homogeneous medium is t = d c. = nd c0. , where nd is optical path length. Fermat's principle: An optical rays always chooses ...
  17. [17]
    [PDF] 1 Light Rays - Wiley-VCH
    The contribution of a path element ds to the optical path length is d opt = n ds = net ⋅ dr, where et = dr∕ds is the tangential unit vector of the ...<|control11|><|separator|>
  18. [18]
    [PDF] ray optics - 1.1
    Hint: In accordance with Fermat's principle the optical path lengths between the two points must be equal for all paths. Lenses. A spherical lens is bounded by ...
  19. [19]
    Tracing rays through graded-index media: a new method
    We present here a new numerical method for tracing rays through graded-index media and show that this method requires much less computational effort.
  20. [20]
    Ray Tracing Method of Gradient Refractive Index Medium Based on ...
    Using the algorithm proposed in this paper to trace the gradient refractive index medium from 0 to 2000 mm, it is possible to obtain the optical path length L 1 ...
  21. [21]
    [PDF] 1207.1.K.pdf - Caltech PMA
    Such an expan- sion, in this context of wave propagation, is called the geometric optics approximation or the eikonal approximation (after the Greek word ǫικων ...
  22. [22]
    An Introduction to Mirages - A Green Flash Page
    Mirages are not optical illusions, as many people (and Web sites!) think. They are real phenomena of atmospheric optics, caused by strong ray-bending in layers ...
  23. [23]
    [PDF] Waves and Imaging
    both directions so that the optical path difference is zero for a stable configuration. ... 2= OPD = nd → d = m · λ. 2 in air. 10.3 DIFFRACTION. In ...
  24. [24]
  25. [25]
    [PDF] The Michelson Interferometer - UMD Physics
    by the optical path difference, which we will denote by ∆x = x2 − x1. A related quantity is the phase difference, ∆φ, given by. ∆φ = 2π λ. ∆x = k∆x,. (1).
  26. [26]
    [PDF] Optical Interferometry ET E1 E2 a) b) - MIT
    The objective of this experiment is to observe the interference of light by combining coherent monochromatic light beams using a Michelson interferometer.
  27. [27]
    26 Optics: The Principle of Least Time
    In other words, the principle is that light takes a path such that there are many other paths nearby which take almost exactly the same time. The following ...<|control11|><|separator|>
  28. [28]
    Fermat's principle - ISB WWW2
    The historical form of Fermat's principle (The path between two points taken by a beam of light is the one which is traversed in the least time.) is incomplete.
  29. [29]
    3.5: Fermat's Principle - Physics LibreTexts
    Nov 7, 2024 · This principle states (in its simplest form) that light waves of a given frequency traverse the path between two points which takes the least time.
  30. [30]
    [PDF] Fermat's Principle and the Geometric Mechanics of Ray Optics
    Figure 4: Fermat's principle states that the ray path from an observer at A to point B in space is a stationary path of optical length. For example, along a ...
  31. [31]
    Reflection and Fermat's Principle - HyperPhysics Concepts
    Fermat's Principle: Light follows the path of least time. Snell's Law can be derived from this by setting the derivative of the time =0. We make use of the ...<|separator|>
  32. [32]
    [PDF] TIE-29: Refractive Index and Dispersion
    Refractive index (n) is the ratio of light speed in vacuum to a medium, measured relative to air. Dispersion is the change of refractive index with wavelength.Missing: physics | Show results with:physics
  33. [33]
    [PDF] Physics of Light and Optics
    ... blue light to be greater than that for red light? Make a sketch of n as a ... optical path length (OPL) between points A and B: OPL|. B. A ≡. B. Z. A. ndℓ.
  34. [34]
    [PDF] SIMG-455 Physical Optics SPSP-455 Optical Physics
    Apr 29, 2008 · 2.0.3 Optical Path Length . ... which means that the phase of the transmitted light increases with increasing radial distance from.
  35. [35]
    31 The Origin of the Refractive Index - Feynman Lectures
    Such a negative slope is often referred to as “anomalous” (meaning abnormal) dispersion, because it seemed unusual when it was first observed, long before ...
  36. [36]
    [PDF] The Scattering Error Correction of Reflecting-Tube Absorption Meters
    ... absorption bands 12. Anomalous dispersion near absorption bands cause changes in the real part of the index of refraction and thus the volume scattering ...
  37. [37]
    [PDF] Phase shift effects in Fabry-Perot interferometry
    The condiLion for consLrucLive inLerference in transmission is that the optical paLh difference be- tween two successive rays should be an integral number of \ ...
  38. [38]
    [PDF] FABRY-PEROT INTERFEROMETER
    If the additional optical path length traveled by a (multiply) reflected beam is an integer multiple of the light's wavelength, then the reflected beams will ...
  39. [39]
    Dynamic Strain Measured by Mach-Zehnder Interferometric Optical ...
    The difference in the optical path induces a relative phase shift in the Mach-Zehnder interferometer. The two optical signals are guided into two of the three ...
  40. [40]
    [PDF] How does a Mach–Zehnder interferometer work? - cs.Princeton
    In fact t is the optical path length through the beamsplitter, which takes account of the actual distance travelled (the beam passes through at an angle) ...
  41. [41]
    [PDF] The Michelson Interferometer - CSUB
    The formula is ∆n = N λ / (2d) where λ is wavelength of the light in vacuum (calculated in the previous part) and d is the length of the vacuum cell (3.0 cm).
  42. [42]
    [PDF] On the Relative Motion of the Earth and the Luminiferous Ether (with ...
    Michelson and Morley-Motion of the Earth, etc. n'-1 n' where. On the undulatory theory, according to Fresnel, first, the ether is supposed to be at rest ...
  43. [43]
    [PDF] 2.2 The Michelson-Morley Experiment
    Jun 24, 2019 · This ether could be used to determine an absolute reference frame (with the help of observing how light propagates through the ether). Note. The ...
  44. [44]
    Mapping optical path length and image enhancement using ...
    We describe the principles of using orientation-independent differential interference contrast (OI-DIC) microscopy for mapping optical path length (OPL).
  45. [45]
    Chapter 1 - Geometrical Optics - SPIE Digital Library
    The technique to trace an optical ray through various optical media is called “sequential raytracing” and is the main way to study geometrical optics. This ...
  46. [46]
    Designing aplanatic thick lenses - SPIE
    Aug 7, 2006 · To be free from spherical aberrations, the optical path length of an axial ray and the optical path length traveled by a marginal ray must be ...
  47. [47]
    Telecentric broadband objective lenses for optical coherence ...
    Surface I indicates schematically the surface at which the central rays from each ray bundle across scans achieve equal optical path lengths (not to scale).
  48. [48]
    [PDF] Controlling light propagation in multimode fibers for imaging ...
    Jun 30, 2023 · In a MMF, different fiber modes have varying group delays. The modal dispersion will broaden the path-length distribution of transmitted ...
  49. [49]
    Index of Refraction - HyperPhysics
    Typical crown glass. 1.52. Crown glasses. 1.52-1.62. Spectacle crown, C-1. 1.523. Sodium chloride. 1.54. Polystyrene. 1.55-1.59. Carbon disulfide. 1.63. Flint ...
  50. [50]
    8.7: Double-Slit Interference - Physics LibreTexts
    Jan 20, 2023 · In 1678 Huygens proposed that each point along a moving wavefront can be represented by a point source of a spherical waves. The wavefronts that ...
  51. [51]
    Influence of Geometric Length Perturbation of Optical Path ... - MDPI
    The results show that the errors introduced by optical path perturbations are linearly related to the magnitude of the perturbations, and the error increases as ...