Fact-checked by Grok 2 weeks ago

Total internal reflection

Total internal reflection is an in which a traveling within a medium of higher encounters the with a medium of lower at an angle of incidence greater than the , resulting in the complete reflection of the back into the original medium without any transmission across the boundary. This occurs because, according to , the refracted would need to bend beyond 90 degrees from the normal, which is impossible, leading to 100% reflection. The , denoted as \theta_c, is the specific angle of incidence at which the refracted ray travels exactly along the boundary, and it is given by the formula \theta_c = \sin^{-1}(n_2 / n_1), where n_1 is the of the denser medium and n_2 is that of the rarer medium (n_1 > n_2). For angles less than \theta_c, partial and occur, but beyond it, total internal reflection takes place, with the reflected ray obeying the law of (angle of incidence equals angle of ). A classic example is in (n_1 \approx 1.33) to air (n_2 = 1.00), where \theta_c \approx 48.6^\circ. Total internal reflection underpins numerous practical applications in and . In optical fibers, light signals are confined within thin or cores by repeated total internal reflections at the core-cladding interface, enabling high-speed data transmission over long distances in and internet infrastructure. Similarly, it is essential for medical endoscopes, where flexible fiber bundles allow internal of the body without invasive . In , the low in diamonds (\theta_c \approx 24.4^\circ) due to their high (n \approx 2.42) causes multiple internal reflections, enhancing their sparkle and brilliance. Additionally, total internal reflection is utilized in prisms for and in optical instruments.

Basic Principles

Optical Description

Total internal reflection (TIR) is the in which a incident on the boundary between two media is completely reflected back into the originating medium, with no into the second medium. This occurs specifically when travels from a medium with a higher (n_1) to one with a lower (n_2 < n_1), and the angle of incidence—measured from the normal to the interface—exceeds a threshold known as the critical angle./University_Physics_III_-Optics_and_Modern_Physics(OpenStax)/01%3A_The_Nature_of_Light/1.05%3A_Total_Internal_Reflection) TIR is a fundamental aspect of ray optics, demonstrating the behavior of at interfaces under conditions where refraction cannot occur. The condition for TIR arises from Snell's law, which governs at the boundary: n_1 \sin \theta_1 = n_2 \sin \theta_2, where \theta_1 is the angle of incidence and \theta_2 is the angle of refraction. When \theta_1 is large enough that the equation would require \sin \theta_2 > 1, no real value for \theta_2 is possible since the sine of an angle cannot exceed 1; consequently, the light cannot refract and is instead totally reflected. This total reflection follows the standard law of reflection, with the angle of reflection equal to the angle of incidence. In a geometric depicting TIR, an incident originates in the denser medium and strikes the planar interface at an greater than the , resulting in a reflected that returns into the same medium at an equal to , while no refracted emerges into the rarer medium. This visual representation underscores the complete redirection of energy back into the first medium, contrasting with typical where part of the transmits across the boundary./University_Physics_III_-Optics_and_Modern_Physics(OpenStax)/01%3A_The_Nature_of_Light/1.05%3A_Total_Internal_Reflection) TIR thus represents a limiting case of , building on the basic principles of bending at interfaces.

Critical Angle

The critical angle, denoted as \theta_c, is defined as the angle of incidence in the optically denser medium ( n_1) at which the angle of refraction in the optically rarer medium ( n_2 < n_1) equals 90 degrees, resulting in grazing emergence along the interface. This angle is derived from Snell's law, n_1 \sin \theta_1 = n_2 \sin \theta_2, by setting \theta_2 = 90^\circ, where \sin 90^\circ = 1, yielding n_1 \sin \theta_c = n_2, or \theta_c = \arcsin(n_2 / n_1). The value of \theta_c depends solely on the ratio of the refractive indices n_2 / n_1 of the two media and, in non-dispersive media where n is independent of wavelength, remains constant regardless of the incident light's wavelength. Variations in \theta_c arise from material properties that determine n, such as composition, as well as environmental factors like temperature and pressure, which alter n through changes in density and molecular structure. For example, at a water-air where n_\text{water} \approx 1.33 and n_\text{air} \approx 1, the critical angle is \theta_c \approx 48.6^\circ. Similarly, for a crown glass-air with n_\text{glass} \approx 1.52, \theta_c \approx 42^\circ.

Advanced Topics

Evanescent Wave

In total internal reflection (TIR), occurring when the angle of incidence θ exceeds the critical angle θ_c, the electromagnetic wave incident from a medium with higher refractive index n_1 onto a medium with lower refractive index n_2 < n_1 does not propagate into the second medium as a traveling wave. Instead, it gives rise to an evanescent wave in the rarer medium (z > 0), characterized by an exponentially decaying field amplitude that penetrates a short beyond the but carries no net across the boundary. The arises from the boundary conditions at the interface, derived using and the continuity of tangential field components. Applying , the parallel component of the wave vector is conserved: k_x = (2π n_1 / λ) sin θ, where λ is the vacuum wavelength. In the second medium, the z-component of the wave vector becomes imaginary for θ > θ_c, yielding a number k_z = (2π n_1 / λ) √(sin² θ - n²), with n = n_2 / n_1 < 1 and the positive real square root ensuring decay. The electric field amplitude in the rarer medium then takes the form E ~ exp(i k_x x - k_z z), representing propagation parallel to the interface (along x) but exponential decay perpendicular to it (along z). The penetration depth δ, defined as the distance over which the field amplitude decays to 1/e of its value at the interface, is given by δ = 1 / k_z = λ / [2π n_1 √(sin² θ - n²)]. This yields values typically on the order of the wavelength λ / (2π n_1), increasing as θ approaches θ_c from above or as the refractive index contrast (1 - n) decreases. For example, in a glass-air interface (n_1 = 1.5, n ≈ 0.67, θ_c ≈ 41.8°), δ is much larger than λ near θ_c and decreases to roughly λ/(2π n_1) ≈ 0.1 λ (∼50 nm for visible λ) at larger angles. The dependence on θ and λ highlights the wave's sensitivity to incidence angle and optical frequency. Energy conservation in TIR is maintained despite the evanescent field's presence, as the time-averaged —the measure of electromagnetic energy flux—has a component parallel to the interface but a zero normal (z-directed) component in the rarer medium. This confirms that, while the field oscillates and stores reactive energy near the boundary, there is no net transmission of power across it, with all incident energy reflected back into the first medium. The wave optics description via the evanescent field resolves an apparent paradox in ray optics, where TIR implies no light whatsoever enters the second medium, yet the full electromagnetic treatment reveals a non-propagating field extension that reconciles complete reflection with the continuity of fields at the boundary.

Frustrated Total Internal Reflection

Frustrated total internal reflection (FTIR) refers to the partial disruption of total internal reflection when the evanescent wave generated at the first interface extends across a thin gap and interacts with a second interface or third medium, allowing some light transmission despite the incidence angle exceeding the critical angle. This phenomenon arises in configurations where the gap thickness d is smaller than the evanescent wave's penetration depth \delta, typically on the order of wavelengths in the gap medium. FTIR was first systematically explored in the 19th century but gained modern theoretical and experimental attention through analyses using Maxwell's equations, highlighting its dependence on the precise geometry of the interfaces. The underlying mechanism resembles quantum mechanical tunneling, where the evanescent field in the lower-index gap medium decays exponentially but overlaps sufficiently with the second interface to couple energy into a propagating wave in the third medium. The transmitted intensity I_t follows I_t \propto \exp(-2 k_z d), with k_z = \frac{2\pi}{\lambda} \sqrt{n_1^2 \sin^2 \theta_i - n_2^2} representing the decay constant perpendicular to the interface, \lambda the wavelength, n_1 and n_2 the refractive indices of the higher- and lower-index media, and \theta_i the incidence angle. Transmission efficiency varies with gap distance d, light polarization (s- or p-polarized), and incidence angle near or above the critical value; as d \to 0, frustration becomes complete, mimicking direct refraction across a single interface with no gap. Unlike standard total internal reflection, where no energy propagates beyond the interface, FTIR permits measurable energy transfer across the boundary, though reflectivity remains near unity for d > \delta. The evanescent wave enables this coupling without violating the conditions for total reflection at an isolated interface. Experimental demonstrations of FTIR commonly employ a prism-prism setup, where two right-angle of high-index (e.g., with n \approx 1.65) are positioned with their hypotenuses facing each other, separated by a controllable air gap or thin . A beam incident on the first prism at an angle greater than the (typically around 42° for glass-air) undergoes partial and , with the transmitted fraction observed on a screen or detector as the gap varies from micrometers to millimeters; reflectivity drops sharply as d decreases below \delta \approx \lambda / (2\pi \sqrt{n_1^2 \sin^2 \theta_i - n_2^2}). Such setups, often using visible sources like He-Ne , quantitatively verify the dependence and distinguish FTIR from conventional by measuring the non-zero in the third medium. These configurations underscore how FTIR modifies total internal while preserving high for thicker gaps.

Phase Shifts

In total internal reflection (TIR), the reflected electric field experiences a polarization-dependent phase shift, defined as the argument of the complex Fresnel reflection coefficients r_s and r_p for s- (perpendicular) and p- (parallel) polarizations, respectively. These coefficients have unit magnitude in TIR for lossless dielectrics, so r_s = e^{i \phi_s} and r_p = e^{i \phi_p}, where \phi_s and \phi_p are the respective phase shifts relative to the incident field. The phase shifts arise from the boundary conditions at the interface and differ markedly between the two polarizations, with both starting at zero at the critical angle and approaching \pi radians at grazing incidence, though \phi_p > \phi_s for all supercritical angles. The explicit forms of these phase shifts, derived from the , are given by \tan\left( \frac{\phi_s}{2} \right) = \frac{ \sqrt{ \sin^2 \theta - n^2 } }{ \cos \theta } for s-polarization and \tan\left( \frac{\phi_p}{2} \right) = \frac{ \sqrt{ \sin^2 \theta - n^2 } }{ n^2 \cos \theta } for p-polarization, where \theta is the angle of incidence and n = n_2 / n_1 < 1 is the refractive index of the rarer medium relative to the denser incident medium. These expressions highlight the sensitivity to polarization: the p-polarization shift is smaller due to the n^2 factor in the denominator, reflecting the differing field components parallel and perpendicular to the interface. The physical origin of these shifts lies in the evanescent wave's phase progression along the boundary, which ensures continuity of the tangential field components; this also manifests as the Goos-Hänchen shift, a small lateral displacement of the reflected beam. The phase difference \Delta \phi = \phi_p - \phi_s increases monotonically with \theta beyond the critical angle \theta_c = \sin^{-1} n and reaches a maximum of approximately \pi/2 radians near grazing angles, enabling precise measurements via ellipsometry or interferometry for applications like refractive index determination. This difference is exploited in devices such as the Fresnel rhomb, where two TIR events at 54.6° produce a total \Delta \phi = \pi/2, converting linear to circular polarization. In contrast to external reflection—where phase changes are discrete (0 or \pi radians) and absent at normal incidence—TIR induces continuous, non-zero shifts for both polarizations even at grazing incidence, due to the evanescent nature of the transmitted field.

Examples and Applications

Everyday Examples

Total internal reflection plays a role in the brilliance of diamonds, where the gem's high refractive index of approximately 2.42 results in a small critical angle of about 24° at the diamond-air interface. When light enters a cut diamond, it undergoes multiple internal reflections off the facets before escaping, trapping and redirecting rays to create the sparkling effect observed from various angles. This phenomenon enhances the diamond's fire and scintillation, making it appear more luminous than gems with lower refractive indices. In aquatic environments, total internal reflection contributes to optical illusions involving submerged objects, particularly when light rays from water to air exceed the critical angle of approximately 48.6°. For instance, a swimmer looking upward from underwater sees the above-water world confined to a circular "Snell's window" spanning about 97°, beyond which total internal reflection causes the view to reflect the underwater surroundings, making distant submerged objects or the pool bottom appear invisible or distorted from certain perspectives. This effect explains why fish perceive a limited view of the surface and why the depths of clear pools seem shallower than they are. Fiber optic light pipes, commonly found in children's toys such as illuminated wands or decorative bundles, demonstrate total internal reflection in a simple, observable way. Light introduced at one end of the flexible plastic fibers is confined within the core by repeated total internal reflections at the core-cladding boundary, allowing it to travel along curved paths and emerge brightly at the other end without significant loss, even when bent. These toys illustrate how total internal reflection enables light guidance in non-straight trajectories, mimicking the principle used in more advanced devices. The formation of rainbows provides a natural example of total internal reflection within water droplets suspended in the atmosphere. In a secondary rainbow, sunlight entering a droplet undergoes two internal reflections—each a total internal reflection since the incidence angles exceed the critical angle for water-air—before refracting out, resulting in an inverted color sequence compared to the primary rainbow, which involves only one such reflection. This double reflection scatters red light on the inside and violet on the outside, creating the fainter, higher arc often seen above the primary bow.

Technological Applications

Optical fibers rely on total internal reflection (TIR) to guide light signals over long distances with minimal loss, forming the backbone of modern telecommunications networks. The structure consists of a core with a higher refractive index (n_core) surrounded by a cladding with a lower refractive index (n_cladding), ensuring that light rays incident at angles greater than the are reflected back into the core. Multimode fibers, which support multiple propagation paths, are used for shorter distances like local area networks, while single-mode fibers, confining light to a single path, enable high-bandwidth transmission over hundreds of kilometers in long-haul systems. In optical instruments, TIR is exploited in prisms to redirect light beams efficiently without the need for metallic coatings, achieving near-perfect reflection. Right-angle prisms deviate an incoming beam by 90 degrees through a single TIR event at the hypotenuse face, commonly used in periscopes and binoculars to fold optical paths and reduce instrument size. Porro prisms, employing two TIR reflections, invert and erect images in binoculars, providing a wider field of view compared to roof prisms. Attenuated total reflectance (ATR) spectroscopy is a FTIR technique that utilizes the evanescent wave generated during TIR to analyze samples non-destructively. Infrared light is directed through a high-refractive-index crystal (e.g., diamond or germanium) in contact with the sample, where the evanescent field penetrates only a few micrometers into the sample, absorbing characteristic wavelengths without requiring dilution or preparation. This method is widely applied for identifying solids, liquids, and powders in materials science and pharmaceuticals. Total internal reflection fluorescence (TIRF) microscopy selectively illuminates fluorophores within ~100-200 nm of a specimen's surface, enabling high-resolution imaging of cellular processes. Developed in the 1980s, TIRF uses the evanescent wave from TIR at a glass-water interface to excite only surface-bound molecules, minimizing background fluorescence from the bulk solution. In cell biology, it is particularly valuable for studying plasma membrane dynamics, such as protein diffusion and vesicle fusion. Beyond these, TIR underpins several advanced medical and photonic technologies. In endoscopes, bundles of optical fibers transmit illumination and imaging signals via TIR, allowing minimally invasive visualization of internal organs during procedures like gastroscopy.

History

Discovery

The earliest documented explanations of total internal reflection (TIR) emerged around 1300 AD through independent efforts by the Persian polymath Kamal al-Din al-Farisi and the German Dominican friar Theodoric of Freiberg, who both modeled the formation of rainbows in water droplets using this phenomenon. Al-Farisi, in his treatise Tanqih al-Manazir (completed circa 1308), described light rays entering spherical water droplets, undergoing refraction, internal reflection at the droplet's inner surface, and a second refraction upon exit, thereby accounting for the primary rainbow's angular position. Similarly, Theodoric's De iride (written between 1304 and 1310) detailed a comparable geometric path involving one internal reflection for the primary bow and two for the secondary, emphasizing that the reflection within the denser medium was total when rays struck the boundary at sufficiently steep angles. These qualitative models marked the first recognition of TIR as a key optical process in natural refraction scenarios, predating formal wave theory by centuries. Building on earlier refraction studies, Johannes Kepler's 1604 work Ad Vitellionem Paralipomena hinted at internal reflection effects during his investigations of light bending at interfaces, though he did not fully isolate TIR as a distinct limit. Kepler experimented with prisms and water surfaces to explore refraction laws but attributed rainbow colors more to dispersion than to bounded internal reflections, stopping short of a systematic geometric treatment. It was René Descartes who first formalized TIR geometrically in his 1637 publication Les Météores, systematically describing it as essential for both primary and secondary rainbows through ray tracing in idealized spherical droplets. Descartes calculated specific deviation angles—approximately 138° for the primary rainbow and 129° for the secondary rainbow—by assuming total reflection occurs when light attempts to exit the droplet at angles exceeding a certain threshold, without yet deriving a quantitative critical angle formula. In the pre-wave optics era, these discoveries framed TIR through the lens of geometric optics, noting its qualitative boundary behavior where light rays, upon reaching an interface from a denser medium, either fully reflect or partially transmit depending on incidence angle, distinct from ordinary partial reflection at normal incidence. This perspective established TIR as a fundamental optical limit, enabling predictive models of phenomena like rainbows without invoking underlying wave mechanisms. Descartes' work, in particular, solidified its role in celestial optics, influencing subsequent geometric analyses.

Rival Explanations by Huygens and Newton

In the late 17th century, Christiaan proposed a wave theory of light in his 1678 manuscript Traité de la Lumière, published posthumously in 1690, which provided an early explanation for (TIR). According to Huygens, light consists of longitudinal waves propagating through an elastic , with secondary wavelets emanating from each point on a wavefront. When light encounters the boundary between a denser and a rarer medium at an angle greater than the , the secondary wavelets in the rarer medium cannot keep pace with those in the denser medium due to the difference in wave speeds, resulting in a phase mismatch that prevents propagation and causes complete reflection back into the denser medium. This model predicted no energy transmission across the boundary during TIR, aligning with observations but lacking insight into subtle evanescent effects. In contrast, Isaac Newton advanced a corpuscular theory in his 1704 work Opticks, viewing light as streams of tiny particles or corpuscles emitted from sources. Newton explained TIR by analogizing light particles to mechanical bodies: when corpuscles approach the interface from a denser medium at a steep angle, they lack the tangential velocity or "force" needed to penetrate the rarer medium, where the attractive forces at the boundary are insufficient to alter their path adequately, leading to total reflection similar to a projectile rebounding off a surface. He further applied this to refraction laws, positing that corpuscles are accelerated or decelerated at boundaries due to attractions, with TIR highlighting the particle's inertial behavior in optically rarer media. The rivalry between these models was evident in their attempts to unify explanations for refraction and TIR, though differences emerged at interfaces: Huygens invoked elastic wave propagation and phase coherence, while Newton relied on mechanical attractions and particle dynamics. Both theories addressed rainbows by incorporating TIR for internal reflections within water droplets—Huygens through wave interference and multiple refractions, Newton via successive corpuscular bounces and dispersion by particle size—but diverged on color origins, with Newton emphasizing inherent corpuscular properties over Huygens' wavelength dependence. Neither framework fully accounted for evanescent waves near the boundary, yet Huygens' approach foreshadowed modern wave optics by emphasizing constructive interference, while Newton's corpuscular view dominated until the 19th century due to its alignment with emerging mechanics and experimental support for rectilinear propagation.

Contributions of Laplace and Malus

In the early 19th century, Pierre-Simon Laplace advanced the theoretical framework for total internal reflection within the corpuscular theory of light, deriving the precise formula for the critical angle between two isotropic media as \sin \theta_c = n_2 / n_1, where n_1 and n_2 are the refractive indices of the incident and rarer media, respectively. This formulation provided a predictive model for the onset of total internal reflection, building on earlier empirical observations and emphasizing the role of light particle velocities at boundaries. Étienne-Louis Malus, influenced by Laplace's mechanistic approach, made a pivotal discovery in 1808 by demonstrating that light becomes partially plane-polarized upon reflection from a dielectric surface, such as glass, particularly at angles near the . Observing sunlight reflected from windows during the siege of Cádiz through a calcite crystal, Malus quantified this effect and extended it analytically to reflections at steeper incidence angles, including those leading to , where the reflected beam exhibits near-complete polarization perpendicular to the plane of incidence. His work revealed the directional, vector-like properties of light rays, challenging purely scalar models and enabling quantitative studies of polarized reflections in TIR configurations. Malus further developed this into Malus' law in 1809, stating that the transmitted intensity I of polarized light through an analyzer is given by I = I_0 \cos^2 \theta, where I_0 is the initial intensity and \theta is the angle between the polarization planes of the polarizer and analyzer. Applied to TIR, this law facilitated intensity analysis of polarized reflected beams, highlighting variations in reflectivity for orthogonal polarization components and foreshadowing vector-based treatments of boundary interactions. These contributions by Laplace and Malus marked a transition from geometric optics toward incorporating light's transverse properties, laying essential groundwork for later electromagnetic formulations like the Fresnel equations by emphasizing predictive modeling of reflection phenomena and the non-isotropic nature of light at interfaces.

Fresnel's Contributions

Augustin-Jean Fresnel advanced the wave theory of light through his investigations into total internal reflection between 1817 and 1823, particularly by demonstrating polarization-dependent phase shifts that confirmed wave interference phenomena. In 1817, while studying polarization effects, Fresnel observed that plane-polarized light incident at 45° to the plane of incidence appeared partially depolarized after total internal reflection in glass prisms, which he attributed to differential phase delays between the s- and p-polarized components. He submitted preliminary findings on these phase changes to the French Academy of Sciences in 1818, building on earlier theoretical predictions by Laplace and Malus but providing the first experimental validation. Fresnel conducted key interference experiments using setups with total internal reflection prisms to quantify these phase shifts. By arranging prisms to produce interfering beams where one path involved total internal reflection, he observed shifts in interference fringes attributable to the phase difference φ_p - φ_s between p- and s-polarizations, with measurements showing relative shifts up to 90° at specific angles above the critical angle. These results, detailed in his 1823 memoir, directly supported by demonstrating that total internal reflection introduces a non-trivial phase alteration consistent with wave superposition, rather than Newton's corpuscular predictions of simple 180° shifts. The experiments utilized rhomb-shaped prisms designed for two successive total internal reflections at approximately 54°37', achieving predictable phase differences that aligned with his derived formulas. Fresnel extended his reflection equations to total internal reflection by treating the reflection coefficients r_s and r_p as complex numbers for incidence angles θ > θ_c, where the imaginary parts represent shifts φ_s = -arg(r_s) and φ_p = -arg(r_p). This formulation, presented in his 1823 memoir to the Academy and published that year, yielded expressions such as tan(φ_s/2) = √[(n^2 sin^2 θ - 1)/cos^2 θ] for s-polarization (with n the ), enabling quantitative prediction of the difference that drives the Goos-Hänchen shift—the lateral of the reflected beam due to the angular dependence of these phases. These complex coefficients provided a unified mathematical description of across all angles, from normal incidence to beyond the . Fresnel's work on total internal reflection unified the phenomena of , partial , and total under consistent wave boundary conditions at interfaces, resolving longstanding debates in and paving the way for electromagnetic formulations. His boundary condition approach, treating light vibrations transverse to the propagation direction, profoundly influenced James Clerk Maxwell's development of the electromagnetic theory of light in the 1860s, where similar conditions derive the full set of from . The 1823 memoir thus marked a pivotal in favor of the wave theory, with lasting impact on optical and principles.

References

  1. [1]
    Total Internal Reflection - HyperPhysics
    When light is incident upon a medium of lesser index of refraction, the ray is bent away from the normal, so the exit angle is greater than the incident angle.Missing: definition | Show results with:definition
  2. [2]
    Total Internal Reflection – University Physics Volume 3
    Fiber optics is one application of total internal reflection that is in wide use. In communications, it is used to transmit telephone, internet, and cable ...Total Internal Reflection · Learning Objectives · Fiber Optics: Endoscopes To...
  3. [3]
  4. [4]
    Total Internal Reflection | Physics - Lumen Learning
    (c) Total internal reflection occurs when the incident angle is greater than the critical angle. Snell's law states the relationship between angles and indices ...<|control11|><|separator|>
  5. [5]
    Refraction, Snell's law, and total internal reflection
    Mar 20, 1998 · Total internal reflection. The critical angle can be found from Snell's law, putting in an angle of 90° for the angle of the refracted ray.
  6. [6]
    25.4 Total Internal Reflection – College Physics - UCF Pressbooks
    The incident angle that produces an angle of refraction of 90 ∘ is called critical angle. Total internal reflection is a phenomenon that occurs at the boundary ...
  7. [7]
    Demonstrating Total Internal Reflection - Stony Brook University
    The critical angle can be determined using Snell's Law. The critical angle is equivalent to the angle of incidence at which the angle of refraction is 90 ...
  8. [8]
    [PDF] 24-2 Total Internal Reflection - Physics
    Apply Snell's Law to this situation to derive an equation for the critical angle, θc . Applying Snell's Law, and recognizing that sin(90°) = 1, gives: . .
  9. [9]
    [PDF] Chapter 26 Geometrical Optics - CSUN
    Derive the Snell's law (refraction law):. F th d. t i l f th. f t. 1. 1. 2. 2 ... Critical Angle for Total Internal Reflection, θ c g. , c. )12. 26( sin. 2. −. =.<|control11|><|separator|>
  10. [10]
    [PDF] Lecture 15: Refraction and Reflection
    For incident angles larger than the critical angle, there is no refraction: all the light is reflected. We call this situation total internal reflection. Total ...
  11. [11]
    Index of Refraction - Physics
    Index of Refraction. When light, or any EM wave, travels in a vacuum its ... Air, 1.000293. Water, 1.333. Crown Glass, 1.52. Salt, 1.544. Diamond, 2.419. The ...
  12. [12]
    Abbe's Refractometer (Theory) : Modern Physics Virtual Lab
    Temperature influences the refractive index of a medium primarily because of the accompanying change in density. For many liquids, the temperature coefficient ...
  13. [13]
    [PDF] Snell's law - NYU Physics department
    Keep in mind that the index of refraction of any material varies on temperature, pressure, and the wavelength of the light source. Typically light within ...
  14. [14]
    Evanescent Waves - RP Photonics
    Evanescent waves have rapidly decaying amplitudes without energy transport. They are used in various optical applications.
  15. [15]
    [PDF] 6.3 Total Internal Reflection - BYU
    The wave decays in z, but phase progression (propagation) occurs in x. We call such a wave an evanescent wave. At θi = θc, cosθt = 0 so α = 0 (no decay). |Γ⊥| ...
  16. [16]
    Total Internal Reflection - Richard Fitzpatrick
    Hence, when the angle of incidence exceeds the critical angle, the coefficient of reflection is unity, and the coefficient of transmission zero.
  17. [17]
    Frustrated total internal reflection: A demonstration and review
    Jul 1, 1986 · We review its history and applications in modern optics. A simple theoretical description of the phenomenon is presented using Maxwell's ...
  18. [18]
    Frustrated Total Internal Reflection
    How it works: A 45°-90° prism will deflect a beam of light by total internal reflection. When two such prisms are sandwiched back-to- ...Missing: kappa d) k_z
  19. [19]
    Quantitative investigation of Fresnel reflection coefficients by ...
    Mar 1, 2012 · For total internal reflection, the phase shift can assume values other than 0 and π. The Fresnel rhomb19 utilizes this phenomenon to convert ...
  20. [20]
    The world above the water, as seen by a fish - UCLA ePhysics
    When the incident angle is greater than 48 degrees, all the light is reflected back into the water (total internal reflection). As your pet goldfish in the ...Missing: optical | Show results with:optical
  21. [21]
    Science Outreach
    Description: These fiber optic devices show how we can exploit total internal reflection to transfer light across great distances, and even in non linear ...
  22. [22]
    rainbow_explication.html - UNLV Physics
    The primary rainbow has one total internal reflection and the secondary rainbow has two.
  23. [23]
    25.4: Total Internal Reflection - Physics LibreTexts
    Feb 20, 2022 · Since no light gets into the cladding (there is total internal reflection back into the core), none can be transmitted between clad fibers that ...
  24. [24]
    Optical Fiber Core - an overview | ScienceDirect Topics
    Light travels through the core with little energy loss, because of the total reflection of the light inside the core (Fig. 20.15).
  25. [25]
    Right-Angle Prisms - Thorlabs
    Due to total internal reflection (TIR), Video 1.2 demonstrates how the right-angle prism can be used as a 90° reflector. When the input light is incident on ...
  26. [26]
    Right-Angle Prisms - Evident Scientific
    Often termed a Porro prism, the light beam in this configuration undergoes two internal reflections after it enters the prism and is deviated by 180 degrees ...
  27. [27]
    Attenuated Total Reflectance (ATR) - Bruker
    ATR is an extremely versatile measurement technique for FT-IR spectroscopy. It's easy to use, easy to maintain, and has plenty of options so nearly any sample ...
  28. [28]
    Attenuated Total Reflection - an overview | ScienceDirect Topics
    An ATR accessory provides an IR spectrum of a sample because the radiation at the reflection point probes the sample with an “evanescent wave.” At a frequency ...
  29. [29]
    Total Internal Reflection Fluorescence (TIRF) Microscopy
    Total internal reflection fluorescence microscopy (TIRFM) exploits the unique properties of an induced evanescent wave or field in a limited specimen region.
  30. [30]
    Imaging with total internal reflection fluorescence microscopy ... - NIH
    The advantages of using TIRF include the ability to obtain high-contrast images of fluorophores near the plasma membrane, very low background from the bulk of ...
  31. [31]
    Optical Principles of the Endoscope | Abdominal Key
    Jul 29, 2016 · The transmission of light through a glass fiber depends on the phenomenon of total internal reflection. If a fiber is straight or curved ...
  32. [32]
    Supermirrors - RP Photonics
    An optical supermirror is a Bragg mirror (typically a dielectric mirror) that is optimized for an extremely high reflectance – in extreme cases, larger than 99 ...
  33. [33]
    Photonic Bandgap - an overview | ScienceDirect Topics
    ... total internal reflection to confine the light to the core. Instead, it uses the periodic structure of the holes. The structure creates a photonic bandgap ...
  34. [34]
    Kamal Al-Din Al-Farisi's Explanation of the Rainbow - Muslim Heritage
    May 23, 2008 · This optical phenomenon was explained simultaneously but independently by two scientists, Kamal al-Din al-Farisi and Theodoric of Freiberg.Missing: total internal
  35. [35]
    Theodoric of Freiberg and Kamal al-Din al-Farisi Independently ...
    His observations indicated that the primary rainbow was formed from light rays that had been refracted twice and reflected once. An initial refraction took ...
  36. [36]
    [PDF] Kepler's Explanation of the Rainbow* - CARL B. BOYER
    Kepler here seems to use the term re- fraction in the Aristotelian sense of reflection. Of the former class, the first color is the light of white heat itself, ...Missing: René total<|control11|><|separator|>
  37. [37]
    [PDF] The Mathematics of Rainbows and Caustics - Fall 2021 - R. L. Herman
    Oct 25, 2021 · • René Descartes (1596-1650). Rediscovered - internal reflections. • Pierre de Fermat (1607-1665). Theory of refraction - Snell's Law. • Isaac ...
  38. [38]
    The Project Gutenberg eBook of Treatise on Light, by Christiaan ...
    If Huygens had no conception of transverse vibrations, of the principle of interference, or of the existence of the ordered sequence of waves in trains, he ...Missing: total | Show results with:total
  39. [39]
    The Project Gutenberg eBook of Opticks:, by Sir Isaac Newton, Knt.
    This new Edition of Sir Isaac Newton's Opticks is carefully printed from the Third Edition, as it was corrected by the Author's own Hand, and left before his ...
  40. [40]
    Light through the ages: Ancient Greece to Maxwell - MacTutor
    He proposed a wave theory, but of course a wave has to travel through a medium so Huygens' model included an all pervading aether which carries the wave. This ...
  41. [41]
    [PDF] arXiv:2208.07759v2 [physics.optics] 17 Aug 2022
    Aug 17, 2022 · The exact formula of the critical angle (between two isotropic media) was established only about one century later by. Pierre-Simon Laplace [4].
  42. [42]
    Generalized total internal reflection at dynamic interfaces
    Mar 13, 2023 · Total internal reflection (TIR) is the optical phenomenon according to which light impinging on an interface between two media is totally ...<|control11|><|separator|>
  43. [43]
    Étienne-Louis Malus | Optics, Polarization, Reflection - Britannica
    French physicist who discovered that light, when reflected, becomes partially plane polarized; ie, its rays vibrate in the same plane.
  44. [44]
    Science, Optics and You - Timeline - Étienne-Louis Malus
    Nov 13, 2015 · In 1808, Malus famously discovered that light could be polarized (a term coined by Malus) by reflection as he observed sunlight reflected ...
  45. [45]
    The intertwined history of polarimetry and ellipsometry - ScienceDirect
    Feb 28, 2011 · Major breakthroughs came in the early 1800s when Malus discovered polarization of light by reflection and his cosine-squared law and Fresnel and ...
  46. [46]
    Etienne Louis Malus - Oxford Reference
    He is remembered for his discovery made in 1808 that light rays may be polarized by reflection. He found this while looking through a crystal of Iceland ...
  47. [47]
    Introduction to Polarized Light | Nikon's MicroscopyU
    ... total internal reflection. As a result, this ray is reflected out of the ... Formula 2 - Malus's Law. I = I(o) • cos2θ. where I is the intensity of light ...
  48. [48]
    The Fresnel Connection: Particles versus Waves - Oxford Academic
    Jun 22, 2023 · Laplace was backing Siméon Denis Poisson (1781–1840), at that time 28 years old and already a considerable force in French science. Poisson was ...
  49. [49]
    Fresnel's original interpretation of complex numbers in 19th century ...
    Apr 1, 2018 · Fresnel. , “. Mémoire sur la loi des modifications que la réflexion imprime à la lumière polarisée. ,” Read in 7 January 1823. In Oeuvres, 1 ...
  50. [50]
    [PDF] Reconsidering the Fresnel-Maxwell Theory Shift: How the Realist ...
    Fresnel's original derivation relied heavily on crucial assumptions about the geometrical configuration of light rays: that the incident, reflected and re-.