Fact-checked by Grok 2 weeks ago

Specular reflection

Specular reflection, also known as regular reflection, is a type of where or other bounce off a smooth surface at a specific angle, producing a mirror-like image, in which the angle of incidence equals the angle of reflection relative to the surface . This phenomenon requires the reflecting surface to be microscopically smooth, with irregularities much smaller than the of the incident (typically less than 1 micrometer for visible light in the 400–700 nanometer range), ensuring that incoming rays remain after reflection. In contrast to , which scatters light in multiple directions from rough surfaces, specular reflection concentrates the reflected rays into a single direction, preserving the and enabling clear images in reflective optical systems like mirrors. The law of reflection, a fundamental principle in , governs this process and holds true regardless of the surface's orientation, as long as a local can be defined at the point of incidence. Specular reflection plays a crucial role in various applications, including the design of reflective , such as mirrors, where it minimizes light loss and maintains image fidelity. It also explains everyday observations, like glare from wet roads during night driving or the sharp reflections in calm bodies of water, which can enhance visibility but also cause visual distractions. In advanced contexts, such as and material science, models of specular reflection incorporate factors like surface and material properties to simulate realistic lighting effects.

Definition and Principles

Core Definition

Specular reflection is the mirror-like reflection of , such as , from a smooth surface, where the reflected remains planar and the angle of reflection equals the angle of incidence, resulting in a clear, undistorted of the source. This phenomenon occurs when incoming bounce off the surface in a coherent, organized manner, maintaining the directionality of the original . Unlike , specular reflection directs the energy into a single, predictable direction, enabling applications like mirrors and optical instruments. A key prerequisite for specular reflection is that the surface roughness must be much smaller than the of the incident , ensuring minimal disruption to the . For visible , with wavelengths around 400–700 nm, this is achieved using highly polished surfaces such as optical flats, which maintain flatness to within a of a wavelength, like λ/4 or better. Specular applies not only to electromagnetic , including visible and radio , but also to other wave types like sound on boundaries and water on calm surfaces. The term "specular" originates from the Latin speculum, meaning "mirror," reflecting its association with polished, reflective surfaces. This concept was first systematically explored and described by the 11th-century polymath Alhazen (Ibn al-Haytham) in his seminal work , where he analyzed reflection principles through experimentation. In a basic ray diagram illustrating specular reflection, an incident ray approaches the surface at an angle θ to (a line perpendicular to the surface at the point of incidence); the reflected ray then departs at the same angle θ on the opposite side of , demonstrating the symmetry of the process. This geometric representation underscores the law of reflection, which governs the behavior and is explored in detail elsewhere.

Distinction from Diffuse Reflection

Diffuse reflection occurs when incident light waves are scattered in numerous directions due to microscopic surface irregularities on the scale of the light's , resulting in no formation of a distinct , as seen in materials like matte paper. In contrast to specular reflection, which involves the coherent redirection of light rays from a smooth surface to produce a mirror-like , diffuse reflection arises from multiple micro-reflections at varied angles on a rough surface, randomizing the outgoing light directions and disrupting . The primary distinction between the two lies in surface microstructure: specular reflection preserves the phase relationships and spatial properties of the incident , enabling precise , while introduces random phase shifts through scattering, leading to a loss of directional specificity. Intermediate cases, such as glossy surfaces, exhibit a combination of both mechanisms, where a dominant specular component is broadened by moderate roughness, producing blurred highlights alongside scattered . This duality allows for varied visual textures in materials like polished wood or painted plastics. The transition between specular and diffuse behavior is governed by the Rayleigh roughness criterion, which quantifies surface smoothness relative to the λ of the incident . A surface produces predominantly specular reflection if the root-mean-square (RMS) roughness σ satisfies σ ≪ λ/8 (for incidence), ensuring differences between reflected rays remain below π/2 radians and constructive in the specular direction. Conversely, when σ ≈ λ, the variations exceed this threshold, causing destructive in the specular path and favoring diffuse . Visually, specular reflection facilitates the creation of clear virtual images, as the organized ray paths mimic the original wavefront, while diffuse reflection yields uniform illumination across observers' views, minimizing glare and hotspots by distributing light evenly without concentrated reflections. This contrast underlies applications in , where specular surfaces are prized for and diffuse ones for non-distracting .

Law of Reflection

Geometric Formulation

In specular reflection, an incident light ray approaches a smooth reflecting surface and strikes it at a specific point of incidence. At this point, the surface normal is the line perpendicular to the tangent plane of the surface. The reflected ray then emanates from the same point, departing at an angle determined by the geometry of the interaction. This configuration forms the basis of the law of reflection, which governs the directional change of the light ray while preserving its specular nature. The angle of incidence, denoted as \theta_i, is defined as the angle between the direction of the incident ray and the surface normal. Similarly, the angle of reflection, \theta_r, is the angle between the direction of the reflected ray and the same normal. The law of reflection states that \theta_i = \theta_r, ensuring that the reflected ray mirrors the incident ray's approach relative to the normal. This equality holds for all points of incidence on an ideally smooth surface, directing the light coherently in a single outgoing direction. The incident ray, reflected ray, and surface normal all lie within a common plane called the . This planar confinement means that the reflection process occurs entirely within this two-dimensional plane, with no deviation out of plane for specular surfaces. As illustrated in the standard ray diagram, a horizontal reflecting surface is shown with a vertical normal at the incidence point, the incident ray approaching at \theta_i to the , and the reflected ray departing at \theta_r = \theta_i, all aligned in the . This geometric arrangement can be understood intuitively through the principles of continuity and . continuity requires that the of the electromagnetic wave remains matched across the reflecting , which is achieved only when the reflected propagates parallel to the incident one, enforcing equal angles. in the specular case directs the reflected energy into a focused along this path, avoiding dissipation that would occur with unequal angles or .

Vector and Mathematical Formulation

The vector formulation of the law of reflection provides a quantitative means to compute the direction of the reflected ray, essential for simulations in and . Assuming unit s for simplicity, the reflected direction \vec{R} is expressed as \vec{R} = \vec{I} - 2 (\vec{I} \cdot \vec{N}) \vec{N}, where \vec{I} is the unit pointing in the of the incident (towards the reflecting surface), \vec{N} is the unit normal to the surface at the point of incidence, and \cdot denotes the . This equation arises from and applies to , with the prerequisite of familiarity with dot products and vector projections. The derivation begins by decomposing the incident \vec{I} into its components parallel and to the \vec{N}. The parallel (normal) component is the (\vec{I} \cdot \vec{N}) \vec{N}, and the (tangential) component is \vec{I} - (\vec{I} \cdot \vec{N}) \vec{N}. Upon reflection, the tangential component remains unchanged, while the normal component reverses , equivalent to subtracting twice the from \vec{I}, yielding the above. This geometric reversal ensures the reflected lies in the and adheres to the kinematic principles of reflection. To verify that this formula satisfies the core law of reflection—equality of incident and reflected angles with the normal—consider the cosines of these angles. With the convention that \vec{I} points toward the surface, the incident angle \theta_i satisfies \cos \theta_i = -\vec{I} \cdot \vec{N} (negative due to the acute angle convention). For the reflected ray, \vec{R} \cdot \vec{N} = [\vec{I} - 2 (\vec{I} \cdot \vec{N}) \vec{N}] \cdot \vec{N} = \vec{I} \cdot \vec{N} - 2 (\vec{I} \cdot \vec{N}) (\vec{N} \cdot \vec{N}) = \vec{I} \cdot \vec{N} - 2 (\vec{I} \cdot \vec{N}) = - \vec{I} \cdot \vec{N}, since |\vec{N}| = 1. Thus, \cos \theta_r = \vec{R} \cdot \vec{N} = \cos \theta_i, confirming \theta_r = \theta_i. In computational contexts like ray tracing, this vector equation enables efficient calculation of reflection paths for rendering realistic scenes. It forms the basis for specular shading in models such as the Phong illumination model, where \vec{R} determines the intensity of highlights by its alignment with the viewer direction, extending the pure directional law to account for material glossiness.

Reflectivity

Fundamental Concepts

Reflectivity, often denoted as ρ or reflectance, is defined as the ratio of the power of electromagnetic radiation reflected by a surface to the power of the incident radiation, resulting in a dimensionless quantity bounded between 0 and 1. In the context of specular reflection, this coefficient specifically quantifies the mirror-like component of the reflected energy, where light is redirected coherently according to the law of reflection. For ideal perfect reflectors, ρ approaches 1, indicating nearly complete reflection without loss, while ρ = 0 corresponds to total absorption or transmission. Reflectivity can be categorized into types such as normal incidence reflectivity, which measures the at perpendicular incidence, and distinctions between total reflectance (encompassing all reflected light) and specular reflectance (the portion directed specularly, excluding diffuse scattering). For non-scattering surfaces, dictates that the sum of reflectivity ρ, α (the fraction absorbed), and τ (the fraction transmitted) equals 1, ensuring no net energy loss or gain at the . Reflectivity is typically measured using spectrophotometers, which illuminate the surface with monochromatic light and detect the reflected intensity to determine wavelength-dependent values ρ(λ), providing spectral profiles across ranges like the . For example, polished silver exhibits a high normal incidence reflectivity of approximately 0.98 in the visible light range, making it a for efficient specular mirrors. Kirchhoff's law of thermal radiation, applicable at , states that for opaque bodies (where τ = 0), the ε equals the α, which in turn equals 1 - ρ, linking reflective properties to thermal behavior. This principle underscores the reciprocity between and for surfaces in thermodynamic balance.

Angular and Material Dependence

The reflectivity in specular reflection exhibits significant dependence on both the angle of incidence and the of the materials involved at the . For light transitioning between two media with refractive indices n_1 and n_2, the provide the precise reflection coefficients, which squared yield the power reflectivity \rho. These equations account for the state of the light and the resulting transmitted angle \theta_t via , n_1 \sin \theta_i = n_2 \sin \theta_t. For at oblique incidence, the perpendicular (s-polarized) reflectivity is given by \rho_s = \left( \frac{n_1 \cos \theta_i - n_2 \cos \theta_t}{n_1 \cos \theta_i + n_2 \cos \theta_t} \right)^2, while the parallel (p-polarized) reflectivity is \rho_p = \left( \frac{n_1 \cos \theta_t - n_2 \cos \theta_i}{n_1 \cos \theta_t + n_2 \cos \theta_i} \right)^2. The overall reflectivity for is then the average, \rho = (\rho_s + \rho_p)/2. These expressions predict that reflectivity generally increases with the angle of incidence, approaching unity at grazing angles for dielectrics, while remaining relatively constant for metals. At normal incidence (\theta_i = 0), the Fresnel equations simplify to a single form independent of polarization: \rho = \left( \frac{n_1 - n_2}{n_1 + n_2} \right)^2. This yields low reflectivity for common dielectric interfaces, such as air-glass (n_1 \approx 1, n_2 \approx 1.5), where \rho \approx 0.04 across visible wavelengths, explaining the partial transparency of glass. Material properties profoundly influence this angular behavior through the refractive index n, which itself varies with wavelength. Dielectrics exhibit low normal-incidence reflectivity that rises sharply toward grazing angles due to the geometric mismatch in wave vectors at the interface. In contrast, metals maintain high, nearly angle-independent reflectivity—often exceeding 90%—owing to their free electrons, which enable strong collective oscillations that efficiently reflect incident electromagnetic waves across a broad angular range. Wavelength dependence further modulates these effects; for instance, metallic reflectivity decreases at longer infrared wavelengths but remains dominant in the visible spectrum./Optical_Properties/Metallic_Reflection) The derive from the boundary conditions imposed by on the tangential components of the electric and magnetic fields at the interface between two isotropic, homogeneous media, ensuring continuity of the fields and their derivatives. These assumptions hold well for smooth, non-absorbing interfaces but overlook , which scatters diffusely, and detailed mechanisms, such as those in conducting media where complex refractive indices are required for full accuracy.

Optical Consequences

Total Internal Reflection

Total internal reflection (TIR) is a phenomenon that occurs when propagating in a medium with a higher (n₁ > n₂) encounters an interface with a medium of lower at an incidence angle θ_i greater than the θ_c, resulting in complete of the incident power back into the denser medium with no transmission across the boundary (ρ = 1). This condition arises from the principles of , where the reflected ray follows the law of , but the absence of a transmitted ray distinguishes TIR from partial at subcritical angles. The is defined as θ_c = arcsin(n₂/n₁), ensuring that all energy is specularly reflected for θ_i > θ_c. The foundation of TIR lies in , which relates the angles of incidence and transmission at the interface: n₁ sin θ_i = n₂ sin θ_t. At the , the transmitted angle θ_t reaches 90°, grazing the interface, such that sin θ_c = n₂/n₁, and for larger θ_i, no real θ_t exists, leading to total reflection. This derivation highlights how the contrast enforces the boundary condition, preventing propagation into the rarer medium. Historically, the underlying principles of and internal reflection were described in the by Willebrord Snell in 1621 and independently by in 1637, who formalized the law that enables the prediction of TIR. During TIR, although no power transmits, a non-propagating evanescent wave forms in the rarer medium, decaying exponentially away from the over a of typically a few , which allows for interactions like in attenuated total spectroscopy. Additionally, the reflected beam experiences a lateral displacement known as the Goos-Hänchen shift, arising from a shift upon that effectively penetrates the slightly before returning, with the shift magnitude depending on the incidence angle and . This shift, first observed experimentally in , provides a measurable consequence of the evanescent field's influence on the process.

Polarization Effects

When undergoes specular reflection at an between two media, the reflected light becomes partially , with the s-polarized component ( perpendicular to the ) reflecting more strongly than the p-polarized component ( parallel to the ). This preferential reflection arises from the angular dependence of the Fresnel reflection coefficients, which favor s-polarization at most incidence angles. A special case occurs at , \theta_B = \arctan\left(\frac{n_2}{n_1}\right), where n_1 and n_2 are the refractive indices of the incident and transmitting media, respectively, and the p-polarized reflectivity \rho_p vanishes completely. At this angle, the reflected light is entirely s-polarized, while the transmitted light is fully p-polarized. This phenomenon derives from the : \rho_p = 0 when the incident angle \theta_i and transmitted angle \theta_t are complementary, satisfying \theta_i + \theta_t = 90^\circ, as the dipole radiation from induced oscillations in the material cancels in the reflection direction for p-polarization. In configurations involving multiple reflections, such as a stack of parallel glass plates oriented at (known as a pile-of-plates polarizer), successive reflections progressively enhance s-polarization in the reflected beam while transmitting predominantly p-polarized light, achieving high degrees of . Phase differences between s- and p-components across these reflections can further result in for certain incident states. From a quantum perspective, the absence of p-polarized reflection at Brewster's angle stems from the fact that induced dipole oscillations in the material, driven by the incident field, do not radiate perpendicularly to their axis, suppressing emission in the reflection direction for dipoles aligned with p-polarization.

Image Formation

In specular reflection from a plane mirror, the image formed is virtual, erect, and the same size as the object, with the image distance behind the mirror equal to the object distance in front. This virtual image appears at the symmetric position relative to the mirror surface, resulting from the extension of reflected rays backward. Additionally, the image exhibits lateral inversion, where left and right are reversed compared to the object. For curved mirrors, the nature of the image depends on whether the mirror is or and the object's position relative to the . In a mirror, which converges , real and inverted images form when the object is placed beyond the , while , erect, and magnified images occur when the object is within the . The f is half the R, given by f = \frac{R}{2}, with f positive for mirrors. mirrors, which diverge , always produce , erect, and diminished images regardless of object position, with f negative. Image location and for both types are determined by the mirror equation: \frac{1}{f} = \frac{1}{o} + \frac{1}{i} where o is the object distance and i is the image distance (positive for real images in concave mirrors, negative for virtual). Image formation in curved mirrors relies on the paraxial approximation, which assumes small angles relative to the optical axis for accurate predictions using the mirror equation. Ray tracing under this approximation uses three principal rays from the object tip: (1) a ray parallel to the principal axis, which reflects through the focal point; (2) a ray passing through the focal point before reflection, which emerges parallel to the axis; and (3) a chief ray directed toward the mirror's center of curvature, which reflects back along the same path. The intersection of these reflected rays (or their backward extensions for virtual images) locates the image. Aberrations limit the quality of images in spherical mirrors. Spherical aberration occurs because paraxial rays focus at a different point than marginal rays, with outer zones of the mirror focusing closer to the surface in concave mirrors, blurring the image. Chromatic aberration is absent in ideal mirrors since reflection does not disperse wavelengths, though slight effects may arise from wavelength-dependent reflectivity of mirror coatings. These aberrations, particularly spherical, can be corrected using aspheric mirrors, which deviate from spherical curvature to ensure all zones focus at the same point. Specular reflection preserves the geometric arrangement of rays, enabling stereopsis and depth perception in virtual images as the binocular disparity matches that of the apparent object position behind the mirror.

Applications and Examples

Everyday Phenomena

Specular reflection is commonly observed in natural settings where smooth surfaces create mirror-like images. Calm water surfaces, such as those on a still lake or pond, act as horizontal mirrors, reflecting the sky, surrounding landscape, and nearby objects with high fidelity due to the flat interface between air and water. Similarly, wet roads after rain exhibit specular reflection by forming a thin, smooth layer of water that mirrors oncoming headlights, creating bright glares that can dazzle drivers and reduce visibility at night. Dew on leaves in the morning often produces small specular glints, as the curved droplets on smooth leaf surfaces reflect sunlight sharply, contributing to the sparkling appearance of foliage. In household environments, specular reflection plays a practical role in daily routines. Plane mirrors used for grooming, such as those in bathrooms, provide clear virtual images of the face and body by reflecting light rays at equal angles to the incident rays from smooth silvered glass surfaces. Polished metal utensils, like stainless steel spoons or knives, demonstrate specular properties when clean and buffed, reflecting tabletop objects or light sources distinctly to aid in tasks like checking appearance or signaling. Glass windows in homes partially reflect incoming light specularly while transmitting the rest, often creating faint mirror images of indoor scenes superimposed on the outdoor view, especially at oblique angles. Atmospheric conditions can bend light paths to produce specular-like effects through rather than direct . Mirages over hot roads or deserts arise from temperature gradients in the air layers near the ground, where cooler air above warmer air acts like a curved refractive medium, creating inverted, elongated images that mimic specular reflections from an illusory surface. Certain adaptations leverage specular or related reflective mechanisms for survival and display. Iridescent feathers in , such as those of hummingbirds or peacocks, produce angle-dependent specular-like reflections through nanostructured barbules that interfere with , generating shifting colors visible during flight or to attract mates or deter rivals. In cats, the —a reflective layer behind the —enhances low-light by retroreflecting unabsorbed back through the photoreceptors, briefly producing a specular glow in their eyes when illuminated at night. Seasonally, specular reflection becomes prominent in winter landscapes. Ice rinks, with their highly polished frozen surfaces, reflect arena lights and surroundings sharply, creating gliding mirror effects that enhance the visual appeal for skaters and spectators. Frost patterns on windows or ground, forming smooth crystalline layers overnight, exhibit specular glints under sunlight, sparkling as the ice facets mirror the environment before sublimating.

Scientific and Technological Uses

Specular reflection forms the foundation of numerous optical instruments designed to manipulate for and measurement. In reflecting telescopes, parabolic mirrors exploit the precise angular dependence of specular reflection to focus parallel incoming rays from distant objects to a single , minimizing aberrations and enabling high-resolution astronomical observations. Similarly, specialized reflective microscope objectives incorporate mirrors with coatings to achieve chromatic aberration-free across broad ranges, where specular reflection from curved mirrors directs efficiently without . Periscopes utilize prisms that leverage —a form of specular reflection at the glass-air interface—to redirect paths by 90 degrees, allowing visibility around obstacles in applications such as and vehicles without the need for external mirrors. In systems and fiber , specular reflection is engineered for efficient light confinement and amplification. mirrors, consisting of multilayer thin-film stacks, achieve reflectivities exceeding 99.9% through constructive of specularly reflected waves at specific wavelengths, serving as essential components in laser resonators to sustain optical feedback. Optical fibers guide light over long distances by relying on at the core-cladding interface, where specular reflection ensures minimal loss and preserves for and sensing. Advanced coatings further tailor specular reflection for performance optimization. Anti-reflective layers, typically quarter-wave films, reduce surface reflectivity to below 1% via destructive of reflected rays, enhancing in lenses and solar cells. Conversely, high-reflectivity coatings on solar sails maximize specular reflection of to generate through photon transfer, with aluminum or layers achieving near-unity reflectance in the for in space missions. Although primarily optical, specular reflection principles extend briefly to acoustics, where smooth surfaces in systems produce mirror-like echoes for underwater mapping and target detection, analogous to optical specular returns. Contemporary advancements include metamaterials and metasurfaces that enable precise control of specular reflection. These engineered structures manipulate phase and amplitude to achieve perfect anomalous reflection or suppress specular components, facilitating applications in and stealth technologies. In systems for environmental mapping, specular reflections from smooth surfaces provide strong returns for accurate distance measurement and , though multibounce effects require specialized algorithms for interpretation.

References

  1. [1]
    Specular vs. Diffuse Reflection - The Physics Classroom
    Reflection off of smooth surfaces such as mirrors or a calm body of water leads to a type of reflection known as specular reflection.
  2. [2]
    Specular Reflection - RP Photonics
    Specular reflection occurs when light reflects at an equal but opposite angle to the incident light, as on mirrors.What is Specular Reflection? · Other Kinds of Reflection...Missing: physics | Show results with:physics
  3. [3]
    Specular and Diffuse Reflection - Evident Scientific
    Specular reflection is defined as light reflected from a smooth surface at a definite angle, and diffuse reflection, which is produced by rough surfaces.
  4. [4]
    Science, Optics, and You: Light and Color - Reflection of Light
    Nov 13, 2015 · The first, defined as light reflected from a smooth surface at a definite angle, was demonstrated in Figure 1 and is called specular reflection.
  5. [5]
    Specular Reflection - an overview | ScienceDirect Topics
    When reflection occurs at an essentially planar interface with roughness that is small in height and extent compared to the wavelength of the impinging light, ...
  6. [6]
    Optical Flats - RP Photonics
    The surface roughness is also usually very low. Application. Optical flats are mainly used as highly flat reference surfaces in interferometers for checking ...Missing: visible | Show results with:visible
  7. [7]
    [PDF] Figure 298 Rayleigh scattering specular reflection diffuse reflection v5
    Specular reflection occurs in only one direction in an organized manner after sound strikes a smooth boundary. As sound strikes a boundary between two media ...
  8. [8]
    Specular - Etymology, Origin & Meaning
    Originating in the 1570s from Latin specularis, meaning "semi-transparent" or "reflective," this word derives from speculum, meaning "a mirror."Missing: reflection | Show results with:reflection
  9. [9]
    Alhazen's problem | OUPblog
    Mar 12, 2015 · He solved multiple problems in reflection, building upon the notion that the angle of incidence equals the angle of reflection, along with the ...
  10. [10]
    Specular and Diffuse Reflection: Interactive Java Tutorial
    Nov 13, 2015 · At the far left boundary of the Surface Roughness slider, the surface becomes totally flat and exhibits specular reflection of all incident ...
  11. [11]
    Passive sensing around the corner using spatial coherence - Nature
    Sep 7, 2018 · In non-line-of-sight conditions, an ideal “specular” reflector such as a mirror preserves most of the light properties, including the wavefront, ...
  12. [12]
    Reflection - RP Photonics
    Specular reflection basically preserves the spatial properties of light (assuming no strong dependence of reflectivity on propagation direction). That allows ...
  13. [13]
    Introduction to Shaders and BRDFs - Scratchapixel
    In summary, we can describe the appearance of many materials as having a diffuse component and a specular or glossy component.
  14. [14]
    [PP1-2-10] The Rayleigh roughness criterion - Living Textbook
    The Rayleight criterion states that a surface can be considered as smooth (mostly reflecting) if the phase difference is less than π/2 radians. As a consequence ...
  15. [15]
    The reflection and refraction of light - Physics
    Jul 27, 1999 · ... reflection from such objects is known as specular reflection. Most objects exhibit diffuse reflection, with light being reflected in all ...
  16. [16]
    26 Optics: The Principle of Least Time - Feynman Lectures - Caltech
    Thus the so-called law of reflection is θi=θr. That is a simple enough proposition, but a more difficult problem is encountered when light goes from one ...
  17. [17]
    [PDF] Lecture 15: Refraction and Reflection
    This is known as Snell's law. The same logic holds for reflected waves: R is the same and λ is the same (since n1 = n2 for a reflection) therefore θ1 = θ2. ...
  18. [18]
    None
    ### Summary of Huygens' Principle for Law of Reflection
  19. [19]
    [PDF] 11.7 Vector Reflections - Contemporary Calculus
    Practice 1: Find the reflection vector R when an incoming vector V = –1,–2 is reflected by the line –x + 2y = 2 (Fig. 4). Give parametric equations for the.
  20. [20]
    Vector reflection at a surface - Sunshine's Homepage
    Reflection in a nutshell:​​ The formula w = v - 2 * (v ∙ n) * n = v - 2 * v∥ can also be understood from a geometrical point of view. As shown in the figure ...
  21. [21]
    Reflectance - an overview | ScienceDirect Topics
    Reflectance is the ratio of reflected to incident light, a dimensionless number, or the fraction of reflected light, often expressed as a percentage.
  22. [22]
    Solid Sample Reflectance Measurements - Shimadzu
    In addition to measuring liquids, UV-VIS spectrophotometers are used to measure the transmittance and reflectance of solid samples.Reflectance Measurement · Relative Diffuse Reflectance...
  23. [23]
    Diffuse Reflectance and Integrating Spheres - JASCO Inc
    There are two types of reflectance: specular and diffuse, where the sum of the two components is the total reflectance. Specular reflectance is light reflected ...
  24. [24]
    Transmission, Absorption, and Reflection Measurements
    The law of conservation of energy states that the light transmitted throughany sample, plus the sum of reflections by the sample, results in the factor 1 of the ...
  25. [25]
    Reflectance Spectrophotometer - X-Rite
    Jul 27, 2022 · A reflectance spectrophotometer measures color by flashing light onto a surface and measuring the percentage of spectral reflectance of ...
  26. [26]
  27. [27]
    Radiation Basics: Making Sense of Emissivity & Absorptivity
    Mar 18, 2025 · Kirchoff's Law states that the total hemispherical emissivity (ε) and total hemispherical absorptivity (α) are equal for surfaces in an ...
  28. [28]
    [PDF] CHAPTER 3 ABSORPTION, EMISSION, REFLECTION, AND ...
    Emissivity measures how strongly a body radiates. For opaque surfaces, absorption and reflection account for all incident radiation, with aλ + rλ = 1.
  29. [29]
    Fresnel Equations - RP Photonics
    Fresnel equations are equations for the amplitude coefficients of transmission and reflection at the interface between two transparent homogeneous media.
  30. [30]
    8.2 Specular Reflection and Transmission
    The Fresnel equations describe the amount of light reflected from a surface; they are the solution to Maxwell's equations at smooth surfaces. Given the index of ...
  31. [31]
    Fresnel Reflectance - Program of Computer Graphics
    Sep 5, 2011 · This decreases the reflectance at normal incidence to 0.04, which is correct for a refractive index of about 1.5, typical of dielectrics such as ...
  32. [32]
    Metal-coated Mirrors - RP Photonics
    Further, the angular dependence of the reflectivity is relatively weak. Metal mirrors play a special role in infrared optics, i.e., in spectral regions where ...
  33. [33]
    Fresnel Equations - The University of Arizona
    The authoritative reference for optics is: M.Born and E.Wolf, "Principles of Optics", 7th ed. Cambridge University Press, 2019. (ISBN: 978-1-108-47743-7) ...Missing: source | Show results with:source
  34. [34]
    Total Internal Reflection – University Physics Volume 3
    (c) Total internal reflection occurs when the incident angle is greater than the critical angle. In figure a, an incident ray at an angle theta 1 with a ...
  35. [35]
    Total Internal Reflection - HyperPhysics
    The critical angle can be calculated from Snell's law by setting the refraction angle equal to 90°. Total internal reflection is important in fiber optics and ...Missing: condition | Show results with:condition
  36. [36]
    [PDF] The Physics of Light Transport - UCSD CSE
    In 1611, Johannes Kepler discovered total internal reflection and described the small angle approximation to the law of refraction. In 1621, Willebrord Snell ...
  37. [37]
    Total Internal Reflection - Richard Fitzpatrick
    When total internal reflection takes place, the evanescent transmitted wave penetrates a few wavelengths into the lower refractive index medium. The ...
  38. [38]
    Goos-Hanchen shift - Jens Nöckel - University of Oregon
    Nov 17, 2012 · The Goos-Hänchen effect is a phenomenon of classical optics in which a light beam reflecting off a surface is spatially shifted as if it had ...
  39. [39]
    The Feynman Lectures on Physics Vol. I Ch. 33: Polarization
    It was discovered empirically by Brewster that light reflected from a surface is completely polarized if the reflected beam and the beam refracted into the ...Missing: specular | Show results with:specular
  40. [40]
    [PDF] Reflection and transmission at oblique incidence - UF Physics
    I'll first discuss the relations amongst the angles of incident, reflected, and transmitted rays and then go on to obtain the transmission and reflection coeffi ...
  41. [41]
    [PDF] 5. Plane Electromagnetic Waves ( ) ( ) ) - Galileo and Einstein
    But an oscillating dipole, as we shall learn, emits zero radiation along its axis. This angle is called Brewster's angle. For unpolarized incident radiation ...Missing: explanation | Show results with:explanation
  42. [42]
    [PDF] Chapter 23 The Reflection of Light: Mirrors
    The Formation of Images by a Plane Mirror. Your image in a flat mirror has four properties: 1. It is upright. 2. It is the same size as you are. 3. The image ...Missing: erect | Show results with:erect<|control11|><|separator|>
  43. [43]
    Physics of Light and Color - Introduction to Mirrors
    Nov 13, 2015 · Finally, when the object rests against the surface of the mirror, the virtual image once again becomes the same size as the object.
  44. [44]
    Mirrors
    The focal length f and the radius of curvature R = 2f are positive for a concave mirror and negative for a convex mirror. For a concave mirror the reflecting ...
  45. [45]
    Images Formed by Curved Mirrors
    The object distance do, the image distance di, and the mirror's focal length f are related by the image equation, Further examples: Consider an object placed ...
  46. [46]
    Mirrors
    Mirrors can form images. In the paraxial approximation we can derive the mirror equation for convex and concave spherical mirrors using the law of reflection.Missing: chief | Show results with:chief
  47. [47]
    Ray Diagrams for Mirrors - HyperPhysics
    Mirror ray tracing uses rays parallel to the optic axis, through the focal point, and through the center of curvature. Back-projecting rays helps construct the ...
  48. [48]
    18.7 Aberration
    A: Mirrors do not have chromatic aberration. A mirrors has only a single surface to grind instead of two for a lens. Mirrors can also be supported from the back ...
  49. [49]
    Perception Lecture Notes: Depth, Size, and Shape
    Stereopsis: greek for "solid sight". Close one eye, and hold up your two index fingers, one fairly close to your face and one as far as you can reach.Missing: specular reflection
  50. [50]
    Light and Color - Reflection of Light - Molecular Expressions
    Nov 13, 2015 · Perhaps the best example of specular reflection, which we encounter on a daily basis, is the mirror image produced by a household mirror that ...
  51. [51]
    Demos: 7A-02 Specular and Diffuse Reflection - Purdue Physics
    But when it is wet it becomes a mirror-like surface and most of the light is reflected forward. Also the oncoming headlights are directly reflected into your ...
  52. [52]
    On the Observation of Unresolved Surface Features of a Planet
    ... leaves as the antisolar point is approached. A stubble field is a uniquely ... Such specular reflection may identify dew drops, however. 2 The diffuse ...
  53. [53]
    Light and Color - Specular and Diffuse Reflection: Interactive Tutorial
    Sep 10, 2018 · Perhaps the best example of specular reflection, which we encounter on a daily basis, is the mirror image produced by a household mirror that ...
  54. [54]
    [PDF] Road surface mirage: A bunch of hot air?
    However, it was also suggested that the mirage is just a phenomenon of specular reflection ... rior mirage phenomenon caused by the temperature gradient.
  55. [55]
    Nanostructural basis of rainbow-like iridescence in common ...
    Angle- resolved specular reflectance of the green region of the feather is shown in Fig. 3, with the incident angle ranging from 10 ° to 45° by 5 ...
  56. [56]
    Feline Vision Problems: A Host of Possible Causes
    Cats also have a specialized layer of tissue beneath the retina that reflects incoming light. This structure—the tapetum lucidum—reflects light not absorbed by ...