Fact-checked by Grok 2 weeks ago

Critical angle

In , the critical angle is defined as the smallest angle of incidence for a ray traveling from a medium with a higher (denser medium) to one with a lower (rarer medium) at which the angle of is exactly 90 degrees, causing the refracted ray to graze along the boundary surface. Beyond this angle, occurs, where all incident is reflected back into the denser medium without any transmission into the rarer medium. This boundary between and is a direct consequence of , which relates the angles of incidence and to the refractive indices of the two media: n_1 \sin \theta_i = n_2 \sin \theta_r, where n_1 > n_2, \theta_i is the angle of incidence, and \theta_r is the angle of . While primarily an optical concept, the term "critical angle" is also used in other fields such as . The critical angle \theta_c is calculated by setting \theta_r = 90^\circ in Snell's law, yielding \sin \theta_c = \frac{n_2}{n_1}, assuming n_1 > n_2; for example, the critical angle for light passing from water (n \approx 1.33) to air (n = 1.00) is approximately 48.6 degrees. This value depends solely on the ratio of the refractive indices and is independent of the wavelength of light in non-dispersive media, though dispersion can cause slight variations in materials like glass. Total internal reflection at or above the critical angle results in 100% reflection efficiency, unlike partial reflection at smaller angles, and it preserves the light's polarization properties. The phenomenon underpins numerous practical applications, particularly in optical technologies. In fiber optics, light signals are confined within thin glass or plastic cores by repeated total internal reflection at the core-cladding interface, where the cladding has a lower , enabling efficient long-distance data transmission in with minimal loss. Similarly, prisms in optical instruments, such as periscopes and , utilize total internal reflection to achieve high-reflectivity mirrors without metallic coatings, providing durable and efficient light redirection. In gemology, the critical angle explains the brilliance of diamonds, as their high (approximately 2.42) results in a small critical angle (about 24 degrees), trapping light internally to enhance sparkle when cut properly. These applications highlight the critical angle's role in advancing fields from communications to .

Definition and Fundamentals

Definition in Optics

In optics, the critical angle, denoted as \theta_c, is defined as the smallest angle of incidence—measured relative to to the —at which propagating from a medium with a higher to one with a lower experiences , such that the refracted propagates along the boundary at an angle of 90 degrees to . This occurs when the no longer refracts into the second medium but is entirely reflected back into the first, marking the threshold beyond which ceases. The basic setup involves a incident from medium 1, characterized by refractive index n_1, onto the interface with medium 2, where n_2 < n_1, such that when the angle of incidence \theta_i equals \theta_c, the angle of refraction \theta_r is precisely 90 degrees, causing the refracted to graze the surface. For angles of incidence greater than \theta_c, total internal reflection takes place, confining the within the denser medium. To understand this, prerequisite concepts include the , which is the angle between the incident ray and the perpendicular (normal) to the interface, and the , which is the corresponding angle of the transmitted ray relative to the normal; these angles determine how light bends at the boundary due to the change in medium speed.

Relation to Snell's Law

, also known as the law of refraction, describes the relationship between the angle of incidence and the angle of refraction when light passes from one medium to another, stated as n_1 \sin \theta_i = n_2 \sin \theta_r, where n_1 and n_2 are the refractive indices of the first and second media, respectively, \theta_i is the angle of incidence measured from the normal, and \theta_r is the angle of refraction. The n is defined as the ratio of the speed of light in vacuum to its speed in the medium, n = c / v, reflecting differences in wave propagation speeds between media. The critical angle emerges directly from Snell's law when the refracted ray grazes the interface, corresponding to \theta_r = 90^\circ, at which point \sin \theta_r = 1 (see the mathematical derivation for the explicit formula). This is applicable only when n_1 > n_2, defining the transition to for larger incident angles. This relation underscores the critical angle as a direct consequence of the speed differences between , where bends away from when entering a rarer medium (n_1 > n_2), allowing the refracted angle to reach 90°; conversely, entry into a denser medium bends toward , preventing such a . The phenomenon occurs exclusively from denser to rarer because only then can the incident produce a refracted ray at or beyond 90°, leading to . The value of the critical angle depends on factors such as due to , where the varies with wavelength—typically increasing for shorter wavelengths—thus altering \theta_c across the spectrum. also influences it, as higher temperatures reduce the by decreasing medium density and altering electronic , generally increasing \theta_c.

Total Internal Reflection

Conditions for Occurrence

Total internal reflection occurs only when propagates from a medium with a higher (n_1) to one with a lower (n_2), satisfying the condition n_1 > n_2. This setup ensures that bends the toward the normal when the angle of incidence is less than the critical angle (\theta_c), but sets the stage for complete reflection beyond it. The angle of incidence must equal or exceed \theta_c, defined as the incidence angle at which the refracted ray would propagate exactly along the (angle of refraction = 90°). At precisely the critical angle, the refracted ray travels parallel to the boundary surface, marking the transition point where any further increase in the incidence angle prevents transmission into the second medium. For incidence angles greater than \theta_c, takes effect, with 100% of the incident reflecting back into the first medium and no transmitting across the . Although no propagating wave enters the rarer medium, a non-propagating exists near the in the second medium, decaying rapidly away from the boundary without carrying forward. These conditions do not hold if n_1 \leq n_2, as light will always refract into the second medium rather than reflect totally, regardless of the incidence angle. Additionally, real-world interfaces may exhibit imperfections such as surface roughness or contaminants, which can scatter light and reduce the efficiency of total internal reflection by allowing partial transmission or absorption. In practical applications, such as optical fibers, these limitations are mitigated by using smooth, clean boundaries or cladding layers to maintain the refractive index contrast.

Physical Mechanism

The physical mechanism of total internal reflection arises from the wave nature of at the between two , where the incident wave from a denser medium (higher refractive index n_1 > n_2) encounters a phase velocity mismatch. When the angle of incidence exceeds the critical angle, the parallel components of the wave vectors must match at the boundary to satisfy conditions, but the perpendicular component in the rarer medium becomes imaginary. This results in a refracted wave that cannot propagate as a traveling wave; instead, it manifests as an evanescent wave, which decays exponentially away from the interface without transporting net into the second medium. From an perspective, the incident wave induces oscillating dipoles at the , which radiate secondary wavelets according to Huygens' principle. For incidence angles greater than the critical angle, these wavelets interfere constructively in the direction of the reflected wave within the denser medium, while in the rarer medium, the mismatch prevents constructive for any propagating transmitted wave. The "refracted" component thus contributes only to a non-propagating , ensuring destructive for energy transmission across the boundary. This pattern reinforces the total reflection, with the evanescent wave serving to maintain field continuity without allowing power flow. Energy conservation is preserved in ideal, lossless media during , as the magnitude of the reaches unity, directing all incident energy back into the first medium with no or . The evanescent wave, while present near the , carries no net Poynting flux perpendicular to the , thus avoiding any energy loss. In ray diagrams, this is visualized as a gradual increase in the reflected 's and a corresponding decrease in the refracted 's as the incidence approaches the ; beyond it, the refracted disappears entirely, leaving only the reflected with a slight lateral shift due to the evanescent field's penetration.

Mathematical Derivation and Calculation

Deriving the Critical Angle Formula

The critical angle \theta_c arises in the context of at an between two media and is derived directly from , which governs the bending of light as it passes from one medium to another. states that for light incident from medium 1 with n_1 to medium 2 with n_2 (where n_1 > n_2), the relationship between the angle of incidence \theta_i and the angle of refraction \theta_r is given by n_1 \sin \theta_i = n_2 \sin \theta_r. This derivation assumes the media are homogeneous and isotropic, the light is monochromatic, and there is no absorption, ensuring plane-wave propagation and a well-defined refractive index. To find the critical angle, consider the limiting case where the refracted ray travels along the interface, such that \theta_r = 90^\circ and \sin \theta_r = 1. Substituting into yields n_1 \sin \theta_c = n_2 \cdot 1, where \theta_c is the critical angle of incidence. Solving for \theta_c gives \theta_c = \arcsin \left( \frac{n_2}{n_1} \right). Here, \theta_c is typically expressed in degrees for practical calculations, though radians may be used in theoretical contexts; for instance, the critical angle at a -air (n_\text{water} = 1.33, n_\text{air} = 1.00) is approximately $48.6^\circ. For incident angles greater than \theta_c, occurs, with no light transmitted into the second medium.

Numerical Examples and Calculations

To illustrate the application of the critical angle formula, consider the air-crown glass interface, where the of crown glass is approximately 1.52 and that of air is 1.00. The critical angle \theta_c is calculated as \theta_c = \sin^{-1}\left(\frac{n_2}{n_1}\right) = \sin^{-1}\left(\frac{1.00}{1.52}\right) \approx \sin^{-1}(0.658) \approx 41.1^\circ, using a standard or computational tool for the inverse sine function. This step-by-step process involves first determining the ratio of the refractive indices, then evaluating the arcsine to obtain the angle in degrees. Another common example is the -air interface, with a of at 1.33, yielding \theta_c = \sin^{-1}\left(\frac{1.00}{1.33}\right) \approx \sin^{-1}(0.752) \approx 48.6^\circ. For the -air interface, where the of is 2.42, the calculation gives \theta_c = \sin^{-1}\left(\frac{1.00}{2.42}\right) \approx \sin^{-1}(0.413) \approx 24.4^\circ. These values demonstrate how a higher in the denser medium results in a smaller critical angle, facilitating at shallower incidence angles. The following table summarizes critical angles for selected common media interfacing with air (n=1.00), based on standard refractive indices at visible wavelengths:
MediumRefractive Index (n)Critical Angle (\theta_c, degrees)
Water1.3348.6
Ice1.3149.8
Crown Glass1.5241.1
Flint Glass1.6238.0
Diamond2.4224.4
Refractive indices sourced from established tables; critical angles computed via \sin^{-1}(1/n). The critical angle can vary slightly with due to , where the refractive index increases for shorter wavelengths (e.g., ) compared to longer ones (e.g., ). For crown , this results in a higher \theta_c for (n ≈ 1.51, \theta_c ≈ 41.5°) than for (n ≈ 1.53, \theta_c ≈ 40.7°), as the lower index for yields a larger arcsine argument. Similar dispersion effects occur in (n_red ≈ 1.331, \theta_c ≈ 48.7°; n_blue ≈ 1.340, \theta_c ≈ 48.2°) and (n_red ≈ 2.407, \theta_c ≈ 24.6°; n_blue ≈ 2.448, \theta_c ≈ 24.0°).

Applications in Technology

Optical Fibers and Waveguides

Optical fibers rely on the principle of total internal reflection (TIR), where the critical angle plays a central role in confining light within the fiber core. The fiber consists of a core material with a higher refractive index n_\text{core} surrounded by a cladding with a lower refractive index n_\text{clad}, ensuring that light rays incident on the core-cladding interface at an angle \theta_i greater than the critical angle \theta_c = \sin^{-1}(n_\text{clad}/n_\text{core}) undergo repeated internal reflections, propagating along the fiber without escaping into the cladding. This structure allows for efficient light guidance over long distances, as demonstrated in early fiber optic designs where the refractive index difference \Delta n = n_\text{core} - n_\text{clad} is typically small, on the order of 1-2%, to optimize the critical angle for practical angles of incidence. The angle, or the maximum angle at which light can enter the endface and still be guided by TIR, is directly tied to the through the (). Defined as \text{NA} = \sin \theta_a = \sqrt{n_\text{[core](/page/Core)}^2 - n_\text{clad}^2}, the NA quantifies the light-gathering capability of the , with typical values ranging from 0.1 to 0.5 for multimode s, enabling efficient from sources like lasers. This relation ensures that only rays within the satisfy the TIR condition inside the , preventing at the core-cladding boundary. Optical fibers are classified into step-index and graded-index types based on their refractive index profiles, which influence how the critical angle affects mode propagation. In step-index fibers, the abrupt change in refractive index at the core-cladding leads to discrete ray paths governed strictly by the critical angle, supporting multiple modes in multimode variants (core diameters ~50-1000 μm) for short-distance applications like data centers. Single-mode fibers, which are usually step-index with core diameters ~8-10 μm, support only the fundamental mode for long-haul transmission. Graded-index fibers, by contrast, feature a gradual decrease in n from core center to cladding, reducing while still relying on TIR at angles exceeding \theta_c, and are used in multimode fibers for applications requiring minimized over moderate distances. Multimode fibers handle higher bandwidths over shorter links, while single-mode fibers achieve lower , enabling transoceanic cables. The advantages of this TIR-based design include exceptionally low signal loss over extended distances, with intrinsic material attenuation as low as 0.2 /km at 1550 nm in silica , making them indispensable for telecommunications infrastructure that spans global networks. However, practical limitations such as losses occur when the causes rays to strike the at angles below \theta_c, leading to losses that are mitigated through optimized cladding designs and minimum bend radii. These properties have driven widespread adoption in high-speed backbones and sensing applications, underscoring the critical angle's role in scalable .

Prisms and Lenses

In optical prisms, total internal reflection (TIR) is exploited to redirect light beams without the need for metallic coatings, relying on the condition where the angle of incidence θ_i exceeds the critical angle θ_c at a glass-air interface. Right-angle prisms, typically made of high-index glass like N-BK7 (refractive index ≈1.52), achieve a 90° deviation by directing light to strike the hypotenuse at θ_i > θ_c (approximately 41.8° for air-glass), resulting in complete reflection and an inverted image. This design is fundamental for beam steering in compact optical systems. Porro prisms extend this principle using two right-angle prisms arranged to produce a 180° deviation, effectively erecting the while folding the twice via TIR at each . In this configuration, incoming light undergoes sequential reflections where θ_i > θ_c ensures efficient turnaround without loss, commonly yielding a right-handed suitable for viewing instruments. The Porro system's ability to reverse both inversion and left-right makes it ideal for extended . These TIR-based prisms find widespread applications in , where Porro designs provide a wider and stereoscopic separation compared to roof prisms; in periscopes for submerged or obstructed viewing; and in camera viewfinders to relay upright images to the . Unlike silvered mirrors, which degrade over time due to oxidation or , TIR prisms maintain 100% reflectivity indefinitely, offering superior durability, reduced alignment sensitivity, and compactness in harsh environments. In lenses, unwanted TIR can occur at internal glass-air interfaces or edges when light rays approach at angles exceeding θ_c, leading to or ghosting in multi-element systems; anti-reflective () coatings mitigate this by reducing Fresnel reflections across interfaces, thereby minimizing the conditions for TIR and enhancing overall to over 99% per surface. For instance, quarter-wave layers with intermediate refractive indices promote destructive of reflected rays, preventing buildup that could trigger TIR in high-angle paths. This is particularly crucial in achromatic lenses for cameras and eyepieces. The critical angle also manifests in underwater optics, where the water-air (n_water ≈1.33) yields θ_c ≈48.6°, creating the "fish-eye effect" or : an observer submerged sees the above-water world compressed into a 97.2° (twice θ_c), beyond which TIR reflects underwater scenes, distorting the apparent . This phenomenon limits visibility for aquatic imaging systems, such as submersible cameras, unless corrected by wide-angle lenses. Historically, TIR in prisms supported early spectrometers by directing calibration light via small reflecting prisms, as in 1888 Queen instruments where a reference source's was superimposed using TIR for accuracy without external mirrors. In telescopes, prisms with glass-air interfaces enabled Fraunhofer's 1814 solar by dispersing light, though TIR reflection emerged in 19th-century designs for precise beam folding in spectrographs.

Other Contexts

Critical Angle of Attack in

In , the critical angle of attack, denoted as \alpha_c, is the angle between the chord line of an and the relative wind at which the coefficient (C_L) achieves its maximum value, initiating phenomenon where abruptly decreases. This angle represents the threshold beyond which the 's ability to generate is compromised due to disruption. For most conventional subsonic used in wings, \alpha_c typically falls within the range of 12° to 20°, though exact values vary based on and operating conditions. The underlying mechanism at \alpha_c involves the separation of airflow from the airfoil's upper surface, driven by dynamics. As the angle of attack rises, the on the suction side intensifies, promoting a from laminar to turbulent flow in the to resist separation. However, exceeding \alpha_c overwhelms this , causing the to detach and form a low-pressure wake, which reduces by up to 50% or more and sharply increases . This separation often begins at the trailing edge and progresses forward, with the severity depending on and flow regime. The critical angle of attack is influenced by multiple factors, including airfoil shape, Reynolds number, and Mach number. Airfoil camber and thickness distribution affect separation onset; for example, thicker or more cambered profiles can sustain higher \alpha_c by delaying boundary layer instability. The Reynolds number, proportional to airspeed and inversely to viscosity, modulates boundary layer thickness—higher values (e.g., above $10^6) promote turbulent flow that adheres better, increasing \alpha_c by 2–5° compared to low-Re conditions. Compressibility at higher Mach numbers (approaching 0.3) introduces shock waves that accelerate separation, lowering \alpha_c. Representative examples include NACA airfoils: the symmetric NACA 0012 stalls at approximately 15°–18° under low-speed, high-Reynolds conditions, while the cambered NACA 4412 reaches \alpha_c around 16° at similar Re. The importance of \alpha_c in aircraft design cannot be overstated, as it directly governs stall margins and flight safety. Designers incorporate \alpha_c data to optimize profiles for desired performance envelopes, integrate stall prevention features like leading-edge slats, and calibrate warning systems that alert pilots when approaching this limit. In operational contexts, maintaining a safe margin below \alpha_c (often 5°–10°) is crucial during high-load maneuvers, preventing loss-of-control incidents that account for a significant portion of accidents. Seminal studies, such as those on high-angle , underscore its role in enhancing across subsonic to regimes.

Angle of Repose in Materials Science

The angle of repose, denoted as φ, represents the steepest angle at which a pile of granular material remains stable against sliding under the influence of gravity, serving as a key indicator of the material's flowability and stability in materials science. This critical angle is mathematically derived from the balance of forces, where φ = arctan(μ), with μ being the coefficient of static friction between the particles. For instance, dry sand typically exhibits an angle of repose around 34°, while gravel shows a higher value of approximately 40°, reflecting differences in interparticle interactions. These values highlight how the angle quantifies the threshold for static equilibrium in unconsolidated granular assemblies. The underlying mechanism involves the equilibrium between the downward gravitational component along the slope and the upward frictional resistance that prevents particle slippage. At the angle of repose, the induced by exactly matches the frictional , resulting in impending motion without actual flow. This stability is influenced by several material properties: larger particle sizes generally increase the angle due to reduced , while irregular shapes enhance and thus elevate φ compared to spherical particles; moisture content can either increase the angle by promoting in low amounts or decrease it at higher levels by reducing through . Density and also play roles, with denser or rougher grains contributing to steeper stable slopes. In practical applications, the angle of repose is essential in for assessing in excavations and embankments, preventing landslides by informing safe incline designs. In mining operations, it guides the safe stacking of piles and conveyor discharge angles to minimize material scatter and ensure structural integrity. Within the , it evaluates powder flow properties for tablet formulation and capsule filling, where angles below 30° indicate excellent flow and higher values signal potential handling issues like arching in hoppers. These uses underscore its role in optimizing processes for granular handling across disciplines. Unlike the optical critical angle, which governs wave and at interfaces, the angle of repose in denotes a static threshold for particulate systems, devoid of wave propagation dynamics.

References

  1. [1]
    Refraction of Light - StatPearls - NCBI Bookshelf
    Jul 17, 2023 · The critical angle is defined as the angle of incidence that creates an angle of refraction of 90 degrees. It is the largest angle of incidence ...
  2. [2]
    201. 25.4 Total Internal Reflection - University of Iowa Pressbooks
    Total internal reflection occurs when the incident angle is greater than the critical angle. Snell's law states the relationship between angles and indices of ...
  3. [3]
    Critical angle for total internal reflection
    To solve for this critical angle of incidence q1, consider Snell's law with sinq2 set equal to unity: This gives a formula for the critical angle qc,. The ...
  4. [4]
    Total internal reflection
    The angle θc for which sinθc = n2/n1 is the critical angle. For angles greater than the critical angle there is no refracted ray. Details of the calculation:
  5. [5]
    [PDF] Snell's Law and Critical Angle
    1/ni = sin i/sin r or ni = sin r/sin i. Figure 2. We define a special angle, the Critical Angle (CA), as the angle of incidence within a gem for which light is ...Missing: optics | Show results with:optics
  6. [6]
    Total internal reflection; and lenses - Physics
    Jul 25, 2000 · The critical angle can be found from Snell's law, putting in an angle of 90° for the angle of the refracted ray. This gives:Missing: definition | Show results with:definition
  7. [7]
    Total Internal Reflection - Richard Fitzpatrick
    Total internal reflection enables light to be transmitted inside thin glass fibers. The light is internally reflected off the sides of the fiber.
  8. [8]
    Total internal reflection
    Total internal reflection can be exploited to make a perfectly reflecting mirror using only glass, with no metal backing. It is possible to use prisms of ...Missing: applications | Show results with:applications
  9. [9]
    Total Internal Reflection: Diamonds & Fiber Optics
    If the angle of incidence is increased beyond the critical angle, the light rays will be totally reflected back into the incident medium.
  10. [10]
    General Terms - Exploring the Science of Light
    Critical angle – The smallest angle of incidence at which total internal reflection occurs. ... Geometric optics – A filed of physics that deals with light ...
  11. [11]
    Refraction, Snell's law, and total internal reflection
    Mar 20, 1998 · The critical angle can be found from Snell's law, putting in an angle of 90° for the angle of the refracted ray. This gives:Missing: formula | Show results with:formula
  12. [12]
    Refraction, Polarisation and Dispersion of Light Waves
    This specific angle is called the critical angleThe angle of incidence above which light cannot pass from a denser to a less dense medium and is completely ...
  13. [13]
    PHY 106: Snells' Law
    This angle is called the angle of incidence. Depending on the situation the refracted ray may bend toward or away from the normal. To determine this bending, we ...
  14. [14]
    Snell's Law -- The Law of Refraction
    The law of refraction is also known as Snell's Law, named for Willobrord Snell, who discovered the law in 1621. ... The critical angle is the first angle for ...
  15. [15]
    Snell's Law - Engineering LibreTexts
    Jul 5, 2021 · Snell's Law, also known as the Law of Refraction, is an equation that relates the angle of the incident light and the angle of the transmitted light at the ...
  16. [16]
    4.2: Reflection, Refraction, and Dispersion - Physics LibreTexts
    Jan 14, 2019 · The critical angle is the angle of incidence above which the total internal reflectance occurs. ... Dispersion, as a general phenomenon, can occur ...
  17. [17]
    Refractive Index - StatPearls - NCBI Bookshelf
    May 20, 2023 · The index of refraction is an important parameter used in optics to determine the angle by which light is reflected and refracted through ...
  18. [18]
  19. [19]
    Total Internal Reflection - Physics Tutorial
    As mentioned above, the critical angle for the water-air boundary is 48.6 degrees. So for angles of incidence greater than 48.6-degrees, TIR occurs. But 48.6 ...
  20. [20]
    Total Internal Reflection - RP Photonics
    For both s and p polarization, the reflectivity becomes 100% (assuming perfect surface quality) above the critical angle, which is in this case 43.6°.<|control11|><|separator|>
  21. [21]
  22. [22]
    Total Internal Reflection - Working Principle and Applications
    May 12, 2014 · Total internal reflection is the reflection of the ... total internal reflection, an evanescent wave does not affect energy conservation.
  23. [23]
    25.3 The Law of Refraction – College Physics
    The changing of a light ray's direction (loosely called bending) when it passes through variations in matter is called refraction.
  24. [24]
    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. Figure: ...
  25. [25]
    Future Scientists: Total Internal Reflection
    The refractive index of air is 1.0 and the refractive index of water is 1.33 so the critical angle for total internal reflection for a water-air interface is ...
  26. [26]
    Index of Refraction - HyperPhysics
    Typical crown glass. 1.52. Crown glasses. 1.52-1.62. Spectacle crown, C-1. 1.523 ... Heavy flint glass. 1.65. Extra dense flint, EDF-3. 1.7200. Methylene iodide ...
  27. [27]
    25.4 Total Internal Reflection – College Physics - UCF Pressbooks
    The same calculation as made here shows that the critical angle for a ray going from water to air is 48.6 ∘ , while that from diamond to air is 24.4 ∘ , and ...
  28. [28]
    Introduction to Optical Prisms
    ### Summary on Total Internal Reflection in Prisms
  29. [29]
    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 ...
  30. [30]
    Anti-Reflection (AR) Coatings
    ### Summary: How Anti-Reflective Coatings Prevent Unwanted Reflections in Lenses
  31. [31]
    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: critical effect
  32. [32]
    Spectrometers
    Student-type spectrometers like this one have been in common use since 1900 ... prism using total internal reflection from a small prism. To use ...
  33. [33]
    Spectroscopy and the Birth of Astrophysics (Cosmology: Tools)
    There were some early forays into spectroscopy before 1850. Joseph Fraunhofer, for example, mounted a prism in front of the objective lens of a small telescope, ...
  34. [34]
    [PDF] Chapter 5: Aerodynamics of Flight - Federal Aviation Administration
    AOA and then rapidly decreases. 20° AOA is therefore the critical angle of attack. The coefficient of drag curve (orange) increases very rapidly from 14 ...
  35. [35]
    Chapter 1. Introduction to Aerodynamics - Pressbooks at Virginia Tech
    1.15 Angle of Attack. We also need to define the way we relate the orientation of the wing to the flow; i,e., the wing or airfoil angle ...
  36. [36]
    [PDF] NACA 0012 Airfoil Section - NASA Technical Reports Server
    The purpose of this paper is to present a com- prehensive data base of the low-speed aerodynamic characteristics of the NACA 0012 airfoil section. In- cluded ...
  37. [37]
    [PDF] Study of Reynolds number effects on the aerodynamics of a ...
    Jul 15, 2021 · The stall angle and the maximum lift coefficient increased with Reynolds number. Once the flow was separated, the separation point moved ...
  38. [38]
    Aerodynamics of Airfoil Sections – Introduction to Aerospace Flight ...
    Pitching moments are defined as positive when they tend to increase the angle of attack ... Can you explain the terms “stall” and “critical angle of attack” in ...
  39. [39]
    [PDF] Aerodynamic Characteristics of Airplanes at High Angles of Attack
    More specifically, as the stall angle of attack is exceeded, the ratio is reduced to less than 50 percent of the value at a = 0° because of combined shielding ...
  40. [40]
    A review on the angle of repose of granular materials - ScienceDirect
    May 1, 2018 · In the literature, the angle of repose is often assumed to be equal to the residual internal friction angle or the constant volume angle in a ...
  41. [41]
    EngArc - L - Angles of Friction - Engineering Archives
    Clearly, the angle of repose is equal to the angle of static friction φ. Motion impending, Fm = Px φ = φs, If the angle of inclination θ ...
  42. [42]
    Angles of Repose - The Engineering ToolBox
    Tipping or dumping angles for common materials like ashes, sand, earth, shingles and more. ; Graphite, flake, 30 - 45 ; Gravel, 40 ; Gypsum, pulverized, 45 ; Hops, ...
  43. [43]
    Granular Materials (all content)
    The angle of repose is the angle between the horizontal surface and the sloping surface of the pile. The tangent of this angle is the slope of repose.
  44. [44]
    An expression for the angle of repose of dry cohesive granular ...
    We present a model that accurately predicts the angle of repose as a function of particle size, both on Earth and under extraterrestrial gravity.
  45. [45]
    Particle shape characterization using angle of repose ...
    Nov 15, 2002 · In this article we examine the use of the angle of repose and the maximum angle of stability of a slope of granular material poured into water, ...
  46. [46]
    Angle of Repose - Material Testing Expert
    Jun 11, 2023 · Understanding the angle of repose helps in designing equipment, such as bins, mixers, and feeders, to ensure efficient and reliable material ...
  47. [47]
    Flow properties of powdery or granular filling substances of ...
    Jul 15, 2022 · The angle of repose is a criterion for the flowability of a powder and a fundamental parameter in bulk material storage. Generally, the angle of ...