Fact-checked by Grok 2 weeks ago

Optics

Optics is the branch of physics that studies the behavior and properties of , including how it is generated, propagated, detected, and interacts with matter. This field encompasses the generation and propagation of waves, as well as their , , , and at interfaces between different media. , typically understood as within the but extending to and ranges, serves as the primary subject of investigation, with optics bridging classical wave theory and . The discipline divides into key branches that address different scales and phenomena of light. Geometric optics treats light as rays traveling in straight lines, approximating its behavior for applications involving lenses, mirrors, and systems where wave effects are negligible, such as in the of cameras and telescopes. In contrast, , also known as wave optics, explores light's wave nature to explain complex effects like , , and , which are essential for understanding phenomena such as the colors in bubbles or the operation of gratings. Quantum optics further extends this by examining light at the particle level, focusing on photons and their interactions in processes like laser emission and . Optics has profoundly influenced and , enabling innovations across multiple sectors. It underpins optical instruments like microscopes and spectrometers for scientific research, fiber-optic cables for high-speed communication, and lasers for precision manufacturing, medical procedures such as , and in CDs and DVDs. In and , optics supports advanced systems, technologies, and directed-energy applications. These advancements stem from optics' ability to manipulate for information processing, sensing, and energy transfer, making it integral to fields like and . Historically, optics traces its origins to ancient civilizations, with early contributions from Greek scholars like , who in the 3rd century BCE described the straight-line propagation of in his work Optics, and , who explored reflection and in the CE. Medieval advancements came from (Alhazen) in the 11th century, whose established the for studying and vision, refuting emission theories and confirming that travels from objects to the eye. The field accelerated in the 17th century with René ' wave theory and of , followed by Isaac Newton's particle model and ' wave explanation, culminating in the 19th century's synthesis through James Clerk Maxwell's electromagnetic theory and Thomas Young's double-slit interference experiments. The 20th century brought quantum insights from and the development of lasers in 1960, transforming optics into a cornerstone of modern physics and engineering.

Fundamentals of Light

Nature and Properties of Light

is a form of that exhibits both wave and particle characteristics, a phenomenon known as wave-particle duality. In its wave aspect, propagates as oscillating electric and magnetic fields perpendicular to each other and to the direction of travel. The particle nature is evident in discrete packets of energy called photons, which carry but have no rest mass. As an electromagnetic wave, light is characterized by its , the distance between successive crests, and , the number of oscillations per second, with the two related by the in c = \lambda \nu. The in is exactly $299\,792\,458 m/s, often approximated as $3 \times 10^8 m/s for conceptual purposes. The of a is given by E = h \nu, where h is Planck's constant, h = 6.626 \times 10^{-34} J s, linking the wave directly to quantized energy levels. The portion of the electromagnetic spectrum visible to the human eye spans wavelengths from approximately 400 nm (violet) to 700 nm (red), corresponding to frequencies between about $4.3 \times 10^{14} Hz and $7.5 \times 10^{14} Hz. This narrow range determines color perception, with shorter wavelengths appearing as higher-energy colors like blue and longer ones as lower-energy reds. In homogeneous media, light propagates in straight lines, a principle known as rectilinear propagation, which underpins many optical phenomena and assumptions in ray optics.

Electromagnetic Theory of Light

The electromagnetic theory of light posits that light is a form of electromagnetic radiation, consisting of oscillating electric and magnetic fields that propagate through space as transverse waves. This framework was established by James Clerk Maxwell in 1865, who unified electricity and magnetism into a set of four fundamental equations describing the behavior of electromagnetic fields. These equations predict that disturbances in the electric field generate magnetic fields, and vice versa, leading to self-sustaining waves that travel at the speed of light, thereby identifying light itself as an electromagnetic phenomenon./University_Physics_II_-Thermodynamics_Electricity_and_Magnetism(OpenStax)/16%3A_Electromagnetic_Waves/16.02%3A_Maxwells_Equations_and_Electromagnetic_Waves) Experimental confirmation came in 1887 when Heinrich Hertz generated and detected electromagnetic waves using oscillating electric sparks, demonstrating their propagation and reflection properties akin to light. In , where there are no charges or currents, simplify to two key forms for the \mathbf{E} and \mathbf{B}: \nabla \cdot \mathbf{E} = 0, \quad \nabla \times \mathbf{E} = -\frac{\partial \mathbf{B}}{\partial t} \nabla \cdot \mathbf{B} = 0, \quad \nabla \times \mathbf{B} = \mu_0 \epsilon_0 \frac{\partial \mathbf{E}}{\partial t} Here, \mu_0 is the permeability of free space and \epsilon_0 is the permittivity of free space. The transverse nature arises because the divergence equations imply no sources, so the fields are to the of , and the curl equations ensure \mathbf{E} and \mathbf{B} are mutually . To derive the wave equation, take the curl of the curl equation for \mathbf{E}: \nabla \times (\nabla \times \mathbf{E}) = -\frac{\partial}{\partial t} (\nabla \times \mathbf{B}) = -\mu_0 \epsilon_0 \frac{\partial^2 \mathbf{E}}{\partial t^2} Using the vector identity \nabla \times (\nabla \times \mathbf{E}) = \nabla (\nabla \cdot \mathbf{E}) - \nabla^2 \mathbf{E} and substituting \nabla \cdot \mathbf{E} = 0, this yields the wave equation: \nabla^2 \mathbf{E} = \mu_0 \epsilon_0 \frac{\partial^2 \mathbf{E}}{\partial t^2} For a plane wave propagating in the z-direction, it simplifies to \frac{\partial^2 E}{\partial z^2} = \frac{1}{c^2} \frac{\partial^2 E}{\partial t^2}, where c = 1/\sqrt{\mu_0 \epsilon_0} is the speed of light in vacuum. A similar equation holds for \mathbf{B}. Polarization describes the orientation of the in the plane perpendicular to the propagation direction. In linearly polarized , the electric field oscillates along a fixed straight line, such as horizontal or vertical. occurs when the electric field rotates at constant magnitude in a circle, either clockwise (right-handed) or counterclockwise (left-handed), resulting from two orthogonal linear components of equal amplitude and 90-degree phase difference. is a general case where the components have unequal amplitudes or phase differences other than 90 degrees, tracing an ellipse. is typically unpolarized, a random superposition of all polarizations, while many optical phenomena selectively produce polarized light. Visible light occupies a narrow band in the broader , spanning wavelengths from approximately 400 nm () to 700 nm (), corresponding to frequencies of about 430–750 THz. This region lies between radiation (longer wavelengths, lower frequencies) and (shorter wavelengths, higher frequencies), with the full spectrum extending from radio waves (wavelengths >1 mm) through microwaves, , visible, , X-rays (<10 nm), and gamma rays (<0.01 nm). All these forms obey the same wave equations, differing only in wavelength and frequency, which determine their interactions with matter.

Historical Development

Ancient and Medieval Contributions

Early understandings of optics emerged in ancient Greece through empirical observations and geometric modeling of light's behavior, particularly in reflection. Euclid, in his treatise Optica composed around 300 BCE, treated light rays as straight lines analogous to geometric constructs, formulating the law of reflection where the angle of incidence equals the angle of reflection. This work laid foundational principles for catoptrics, the study of reflected light, by analyzing how rays interact with mirrors to form images. Euclid's approach emphasized mathematical deduction over physical causation, influencing subsequent Greco-Roman and later traditions. Building on these ideas, Claudius Ptolemy advanced the field in the 2nd century CE with systematic experiments on refraction in his Optics. Ptolemy compiled the first quantitative tables of refraction angles for light passing between air and media like water and glass, observing that the refracted ray bends toward the normal when entering a denser medium. These tables, derived from measurements at various incidence angles, approximated but did not precisely capture the modern law of refraction, yet they provided empirical data that persisted into medieval scholarship. Ptolemy integrated refraction into his broader visual ray theory, where sight results from rays emanating from the eye. During the Islamic Golden Age, Ibn al-Haytham (known as Alhazen) revolutionized optics with his comprehensive Book of Optics (Kitāb al-Manāẓir), completed around 1021 CE, marking a shift toward experimental methodology. Rejecting the emission theory of vision—where rays originate from the eye—he established the intromission theory, positing that light rays travel from objects to the eye, enabling accurate perception. Ibn al-Haytham detailed the camera obscura, demonstrating how light forms inverted images through a small aperture in a darkened room, a key insight into image formation without lenses. His work critiqued and refined Ptolemy's refraction tables through controlled experiments, emphasizing causation and quantitative analysis over mere geometry. In medieval Europe, scholars like Witelo and Roger Bacon extended these Islamic and ancient foundations in the 13th century, fostering perspectivist optics that blended geometry with physiology. Witelo's Perspectiva, drawing heavily from , explored refraction's effects on vision, including atmospheric bending and lens interactions, while treating light as propagating rays that the eye receives. Roger Bacon, in his Opus Majus (1267), advocated experimental verification in optics, discussing refraction through media and the magnifying potential of convex lenses to aid presbyopia. Bacon's writings highlighted optics' role in divine order, influencing university curricula. These theoretical advances paralleled practical innovations in optical instruments. Ancient burning mirrors, parabolic devices attributed to figures like Archimedes (3rd century BCE), concentrated sunlight to ignite distant objects, exemplifying early applications of reflection principles. By the late 13th century, rudimentary spectacles emerged in Italy, with convex glass lenses ground by monks in Pisa around 1285 to correct farsightedness, marking the first widespread optical aid. These developments bridged empirical philosophy and utility, paving the way for later scientific rigor.

Scientific Revolution and 19th Century Advances

During the Scientific Revolution in the early 17th century, Johannes Kepler advanced the understanding of refraction and the optics of the eye through his work Dioptrice (1611), where he described the eye as functioning like a camera obscura, with light rays focusing on the retina to form inverted images, and proposed theoretical foundations for lens-based telescopes using convex lenses. Building on this, René Descartes in La Dioptrique (1637) derived the law of refraction using a mechanical analogy of light as particles with tendencies to motion, stating that the ratio of sines of the angles of incidence and refraction equals the ratio of velocities in the two media, which provided a quantitative basis for predicting light bending at interfaces. In the late 17th century, debates over light's nature intensified with Christiaan Huygens' Traité de la Lumière (1690), which proposed a wave theory where light propagates as longitudinal pressure waves in an elastic ether, successfully explaining reflection and refraction via of secondary wavelets. Contrasting this, Isaac Newton in Opticks (1704) advocated a corpuscular theory, positing light as streams of particles with different velocities in media to account for refraction and dispersion, while his experiments on prisms demonstrated that white light decomposes into spectral colors, influencing optical analysis for decades. The 19th century saw the wave theory gain empirical support through Thomas Young's double-slit experiment (1801), in which coherent light passing through two closely spaced slits produced an interference pattern of alternating bright and dark fringes on a screen, providing direct evidence of light's wave superposition and challenging the corpuscular model. Augustin-Jean Fresnel extended this in 1818 with his memoir on diffraction, applying Huygens' principle to predict and explain edge diffraction patterns, including the Poisson spot—a bright central disk in the shadow of a circular obstacle—confirming wave propagation and rectilinear motion as an approximation. Precise measurements of light's speed further solidified its finite velocity and wave-like properties. In 1849, Hippolyte Fizeau used a toothed wheel to interrupt a light beam, measuring its round-trip time over 8.6 km to a mirror, yielding a speed in air of approximately 313,000 km/s, the first accurate terrestrial determination. Léon Foucault refined this in 1850 with a rotating mirror apparatus, confirming light travels slower in water than in air (about 227,000 km/s in water versus 298,000 km/s in air), supporting the wave theory over emission models. The culmination came in 1865 with James Clerk Maxwell's "A Dynamical Theory of the Electromagnetic Field," unifying electricity, magnetism, and light by showing that varying electric and magnetic fields propagate as transverse waves at the speed of light (approximately 299,792 km/s in vacuum), implying light is an electromagnetic phenomenon and predicting radio waves as extensions of the spectrum. This synthesis resolved prior debates and laid the groundwork for modern optics.

Geometrical Optics

Reflection and Refraction Laws

The foundational principles of geometrical optics are encapsulated in the laws of reflection and refraction, which describe how light rays interact with interfaces between media. These laws can be derived from , which states that light travels between two points along the path that requires the least time compared to nearby paths, equivalent to minimizing the optical path length \int n \, ds, where n is the refractive index and ds is the differential path length. This variational principle, proposed by around 1657, provides a unified framework for ray optics and aligns with the wave nature of light by ensuring phase stationarity, though the underlying electromagnetic description is addressed elsewhere. The law of reflection governs the behavior of light at a reflecting surface, stating that the angle of incidence \theta_i equals the angle of reflection \theta_r (\theta_i = \theta_r), measured relative to the normal. To derive this from , consider a ray from point A reflecting off a mirror to point B; the time t for a path reflecting at variable position x is t = \frac{\sqrt{x^2 + h_1^2} + \sqrt{(L - x)^2 + h_2^2}}{c/n}, where c/n is the speed in the medium, h_1 and h_2 are heights, and L is the mirror width. Minimizing t with respect to x yields \sin \theta_i = \sin \theta_r, implying \theta_i = \theta_r for the planar case. Refraction occurs when light passes from one medium to another with different refractive indices n_1 and n_2, bending the ray according to Snell's law: n_1 \sin \theta_1 = n_2 \sin \theta_2, where \theta_1 and \theta_2 are the angles of incidence and refraction, respectively. Historically, this relation was empirically determined by Willebrord Snell in 1621, though first published by René Descartes in 1637. Using Fermat's principle, the time for a ray crossing the interface at variable x is t = \frac{\sqrt{x^2 + h_1^2}}{c/n_1} + \frac{\sqrt{(L - x)^2 + h_2^2}}{c/n_2}; minimization gives n_1 \sin \theta_1 = n_2 \sin \theta_2. The refractive index n is defined as n = c/v, where c is the speed of light in vacuum and v is the speed in the medium. When light travels from a higher-index medium (n_1 > n_2) and \theta_1 exceeds the \theta_c = \sin^{-1}(n_2 / n_1), occurs, with no transmitted ray and full reflection back into the first medium. This condition arises directly from by setting \theta_2 = 90^\circ, so \sin \theta_c = n_2 / n_1. contributes to various optical effects, while mirages arise from due to varying refractive indices in layered media, leading to apparent bending of rays.

Lenses, Mirrors, and Imaging

Lenses and mirrors are fundamental optical elements that manipulate rays according to the laws of and to form images of objects. In , these devices converge or diverge rays to create focused representations, enabling applications from simple magnifiers to complex systems. The formation of images depends on the , material, and positioning of these elements relative to the object and observer. For thin lenses, which approximate lenses where the thickness is negligible compared to the radii of curvature, the relationship between object distance u, image distance v, and f is given by the thin lens equation: \frac{1}{f} = \frac{1}{u} + \frac{1}{v} This equation arises from the paraxial approximation, assuming rays are close to the . Sign conventions follow the Cartesian system: distances to the left of the (object side for a ) are negative for u, positive v indicates a on the opposite side, and f is positive for converging lenses and negative for diverging ones. Spherical mirrors, whether or , follow a similar for under the paraxial : \frac{1}{f} = \frac{1}{u} + \frac{1}{v} Here, the f is half the , positive for mirrors (converging) and negative for (diverging), with sign conventions aligning object distance u as negative when on the incident side. This form mirrors the lens , reflecting the geometric similarity in paths for versus . Images formed by lenses and mirrors can be real or virtual, and upright or inverted, depending on the object's position relative to the . Real images form when rays converge to a point, allowing onto a screen, as in a convex with the object beyond the ; virtual images appear to diverge from a point behind the or mirror, observable only by looking through the element, such as in a . Magnification m quantifies the image size relative to the object, given by m = -\frac{v}{u}, where the negative sign indicates inversion for real images. Despite ideal equations, real optical elements suffer from aberrations that distort images. Spherical aberration occurs because peripheral rays focus at different points than paraxial rays due to the spherical surface geometry, leading to blurred edges. Chromatic aberration arises from the wavelength-dependent refractive index of lens materials, causing different colors to focus at varying distances and producing color fringing. These imperfections highlight the need for corrective designs in precise imaging.

Approximations and Ray Tracing

In geometrical optics, the paraxial approximation simplifies the analysis of optical systems by assuming that light rays make small angles with the , typically less than 10–15 degrees, where the relative remains below 1%. This approximation relies on the small-angle expansions sin θ ≈ θ, tan θ ≈ θ (with θ in radians), and cos θ ≈ 1, which linearize the involved in and calculations. These relations hold because higher-order terms, such as θ³/6 in the sin θ expansion, become negligible for small θ; for instance, at θ = 10°, the in sin θ ≈ θ is about 0.5%, but it exceeds 5% at θ = 30°. The validity is limited to near-axis rays, beyond which aberrations like spherical arise, necessitating more exact methods for wide-angle systems. Ray tracing extends the paraxial framework to model light propagation through multi-element systems by iteratively applying the laws of reflection and refraction at each interface. In sequential ray tracing, a ray's position and direction are tracked from the object through surfaces, using Snell's law (n₁ sin θ₁ = n₂ sin θ₂) at refractive boundaries and the law of reflection (angle of incidence equals angle of reflection) at mirrors; under paraxial conditions, θ ≈ sin θ streamlines computations without significant loss of accuracy for imaging predictions. This method is foundational for lens design software, where rays are traced from multiple object points to assess image quality, though it ignores wave effects like diffraction. For complex systems, such as telescopes, thousands of rays may be traced to map aberrations, with paraxial rays providing initial alignment before exact tracing refines paths. Matrix optics, or the , further simplifies paraxial ray tracing for linear systems by representing each optical element as a 2×2 ABCD that transforms the ray's (r) and (θ) from input to output: \begin{pmatrix} r' \\ \theta' \end{pmatrix} = \begin{pmatrix} A & B \\ C & D \end{pmatrix} \begin{pmatrix} r \\ \theta \end{pmatrix} Here, A, B, C, and D are system-specific coefficients derived from and refractive indices; for example, free-space over distance d yields the \begin{pmatrix} 1 & d \\ 0 & 1 \end{pmatrix}, while thin-lens uses \begin{pmatrix} 1 & 0 \\ -1/f & 1 \end{pmatrix} (f is ). Reflections at curved mirrors follow similar forms, such as \begin{pmatrix} 1 & 0 \\ -2/R & 1 \end{pmatrix} for radius R. The overall system is the product of individual matrices in reverse order, enabling efficient computation of imaging properties like (A) and effective (from det(ABCD) = 1 for lossless systems). This approach, introduced in the mid-20th century, revolutionized optical design by avoiding explicit ray-by-ray calculations for paraxial predictions. For near-collimated beams, such as those from lasers, the approximation builds on paraxial ray tracing by treating the beam as a bundle of rays with a phase profile, characterized by waist size w₀ and range z_R = π w₀² / λ, where is minimal over distances much less than z_R. This links to wave descriptions in modern applications like fiber coupling, where matrices propagate the beam parameters q = z + i z_R.

Physical Optics

Wave Propagation and Superposition

In physical optics, the propagation of light is described using the wave model, where electromagnetic disturbances satisfy the scalar in free space, \nabla^2 E - \frac{1}{c^2} \frac{\partial^2 E}{\partial t^2} = 0, with c as the . A key conceptual tool for understanding this propagation is the Huygens-Fresnel principle, which posits that every point on an existing serves as a source of secondary spherical wavelets that propagate forward at speed c, with the subsequent forming as the tangent envelope to these wavelets, incorporating phase delays based on path lengths. This principle, first articulated by in his 1690 treatise Traité de la Lumière and mathematically formalized by in his 1818 memoir on , enables the reconstruction of wavefront evolution without solving the full directly. Solutions to the wave equation yield fundamental forms for light propagation. Plane waves, representing idealized collimated beams like those from distant sources, take the form E(z, t) = E_0 \cos(kz - \omega t + \phi), where E_0 is the , k = 2\pi / \lambda is the , \omega = 2\pi \nu is the , z is the propagation , t is time, and \phi is a constant; these waves maintain constant amplitude and phase across planes to the of . Spherical waves, approximating emission from point sources such as apertures or scatterers, are given by E(r, t) = \frac{E_0}{r} \cos(kr - \omega t + \phi), where r is the radial distance; the $1/r factor accounts for amplitude diminution due to over expanding spherical surfaces. These solutions assume monochromatic waves in isotropic, homogeneous media and form the basis for more complex propagations via superposition. The linearity of the wave equation implies the principle of superposition, whereby the total electric field at any point is the vector sum of fields from individual waves: E_{\text{total}} = \sum_i E_i./01%3A_Waves_in_One_Dimension/1.04%3A_Superposition_Principle) This allows arbitrary combinations of plane and spherical waves to describe realistic light fields, but stable superposition requires coherence: temporal coherence, quantified by the coherence time or length l_c = c \tau_c, ensures phase predictability over durations \tau_c, while spatial coherence maintains phase relations across transverse extents, often limited by source size via the van Cittert-Zernike theorem. Incoherent superpositions average intensities without phase-dependent effects, whereas coherent ones enable constructive or destructive interference. Wave propagation characteristics are further delineated by and group velocities, tied to the \omega = c k in , where no material occurs. The v_p = \omega / k = c describes the speed of constant-phase surfaces, such as wave crests. The v_g = d\omega / dk = c, representing the propagation of the wave packet envelope and thus the or , equals the phase velocity in non-dispersive media like ; in dispersive media, v_g differs, highlighting how wave packets spread or compress.

Interference and Diffraction

Interference occurs when two or more coherent waves superpose, resulting in regions of enhanced or reduced intensity depending on their phase relationship. This phenomenon provides direct evidence for the wave model of , as first demonstrated in Thomas Young's in 1801. In this setup, monochromatic passes through two closely spaced slits, producing an pattern of alternating bright and dark fringes on a distant screen due to the path length differences between waves from each slit. The spacing between adjacent bright fringes, known as the fringe width Δy, is given by the formula Δy = λL/d, where λ is the of the , L is the distance from the slits to the screen, and d is the separation between the slits./University_Physics_III_-Optics_and_Modern_Physics(OpenStax)/04%3A_Diffraction/4.04%3A_Double-Slit_Diffraction) This relation arises from the condition for constructive , where the path difference is an integer multiple of λ, and the small-angle approximation sin θ ≈ θ = Δy/L. Young's experiment quantitatively confirmed the wave theory by measuring λ from observed fringe patterns, overturning the particle model dominant at the time. Thin-film interference exemplifies how reflections from multiple surfaces within a thin layer produce colorful patterns, such as those seen in soap bubbles or oil slicks. For a thin film surrounded by a lower-index medium, light rays reflecting from the top and bottom experience a path difference of 2nt cos θ, where n is the of the film, t is the film thickness, and θ is the angle of incidence inside the film. The condition for constructive in , accounting for the relative shift of π at one interface, is 2nt cos θ = (m + 1/2)λ for m ≥ 0; however, in or specific configurations without net phase inversion, constructive interference occurs when 2nt cos θ = mλ./University_Physics_III_-Optics_and_Modern_Physics(OpenStax)/03%3A_Interference/3.05%3A_Interference_in_Thin_Films) This leads to wavelength-dependent intensity maxima, explaining the iridescent colors: shorter wavelengths interfere constructively at certain thicknesses, while longer ones may destructively interfere. The effect is prominent when t is on the order of λ, typically hundreds of nanometers, and has applications in anti-reflective coatings where destructive conditions are engineered./University_Physics_III_-Optics_and_Modern_Physics(OpenStax)/03%3A_Interference/3.05%3A_Interference_in_Thin_Films) Diffraction refers to the bending of around obstacles or through , which becomes significant when the size approaches the , deviating from predictions. In single-slit , a passing through a slit of width a produces a central bright maximum flanked by alternating minima and secondary maxima on a screen. The positions of the dark minima are determined by the condition sin θ_m = mλ/a, where m = ±1, ±2, ..., θ_m is the angle from the center, and λ is the ./University_Physics_III_-Optics_and_Modern_Physics(OpenStax)/04%3A_Diffraction/4.03%3A_Single-Slit_Diffraction) This arises from the destructive of wavelets across the slit, as described by Huygens' , with the first minimum at sin θ ≈ θ = λ/a for small angles, giving an angular width of the central maximum of approximately 2λ/a. The pattern follows the , I(θ) = I_0 [sin(β)/β]^2 where β = (π a sin θ)/λ, highlighting the wave nature by spreading beyond the geometric shadow./University_Physics_III_-Optics_and_Modern_Physics(OpenStax)/04%3A_Diffraction/4.03%3A_Single-Slit_Diffraction) These wave effects impose fundamental limits on optical resolution, as quantified by the Rayleigh criterion. For a circular aperture of diameter D, the minimum resolvable angular separation θ_min between two point sources is θ_min ≈ 1.22 λ/D, where the factor 1.22 accounts for the first zero of the Airy diffraction pattern./University_Physics_III_-Optics_and_Modern_Physics(OpenStax)/04%3A_Diffraction/4.02%3A_The_Diffraction_Grating) Introduced by Lord Rayleigh in 1879, this criterion defines resolution as the point where the central maximum of one Airy disk falls on the first minimum of the other, setting the diffraction limit for telescopes and microscopes. For example, in visible light (λ ≈ 550 nm), a 1-m telescope achieves θ_min ≈ 0.07 arcseconds, illustrating how larger apertures enhance resolution by reducing the spread. This limit underscores why physical optics is essential for high-precision imaging, beyond ray-based approximations./University_Physics_III_-Optics_and_Modern_Physics(OpenStax)/04%3A_Diffraction/4.02%3A_The_Diffraction_Grating)

Polarization, Dispersion, and Scattering

refers to the restriction of the oscillations in a to a particular direction perpendicular to the direction of propagation, a consequence of 's nature./University_Physics_III_-Optics_and_Modern_Physics(OpenStax)/01%3A_The_Nature_of_Light/1.08%3A_Polarization) This property arises when interacts with certain materials or surfaces, such as through , leading to partially or fully polarized . In 1809, Étienne-Louis Malus discovered that reflected from a surface at certain angles becomes polarized, laying the foundation for quantitative descriptions of effects. A key relation governing polarized light transmission is Malus's law, which states that the intensity I of polarized light passing 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 direction of the incident and the transmission axis of the analyzer. This law, derived from Malus's experiments with crystals and reflected light, quantifies how the transmitted intensity varies with orientation, enabling precise measurements of states. In anisotropic media, such as certain crystals, causes unpolarized light to split into two rays: the ordinary ray, which follows the standard law of , and the extraordinary ray, which experiences a different depending on the light's direction relative to the crystal's optic axis. This double , first systematically studied in materials like and , results from the medium's directional dependence on the for different polarizations. Dispersion in optics describes the wavelength-dependent variation of the n(\lambda) in a medium, leading to different propagation speeds v = c / n(\lambda) for of varying wavelengths, where c is the in . This phenomenon causes white passing through a to separate into a spectrum of colors, as demonstrated by in his 1704 , producing the familiar pattern due to greater of shorter wavelengths. Materials exhibit when dn/d\lambda < 0, meaning refractive index decreases with increasing wavelength, which is typical for most transparent media in the visible range; conversely, anomalous dispersion occurs near absorption bands where dn/d\lambda > 0, inverting this behavior. Scattering processes further illustrate wavelength-dependent interactions of with . Rayleigh scattering, applicable to particles much smaller than the (such as atmospheric molecules), has an intensity proportional to $1/\lambda^4, scattering shorter wavelengths more efficiently than longer ones, which explains the color of the daytime sky as observed by Lord Rayleigh in his 1871 analysis. This preserves and is isotropic for small particles. For larger particles comparable to or exceeding the , such as aerosols or cloud droplets, dominates, producing less wavelength-selective forward scattering that results in whiter or grayish appearances in or .

Modern Optics

Quantum Nature of Light

The classical description of light as continuous electromagnetic waves, as explored in , successfully accounts for phenomena like and but fails to explain certain observations, such as the of and the behavior of interacting with matter at atomic scales. This limitation prompted the development of , which posits that exhibits a particle-like nature in addition to its wave properties, fundamentally bridging classical and modern optics. A pivotal demonstration of light's quantum character is the photoelectric effect, where light incident on a metal surface ejects electrons, but only if the light's frequency exceeds a material-specific threshold, regardless of intensity. In 1905, Albert Einstein proposed that light consists of discrete energy packets, or quanta, now known as photons, each carrying energy E = h\nu, where h is Planck's constant and \nu is the frequency. The kinetic energy of the ejected electron is then given by KE = h\nu - \phi, with \phi representing the work function, the minimum energy needed to escape the surface; this equation resolved the frequency dependence and intensity's role in providing photon number rather than energy per photon. Einstein's photon concept earned him the 1921 Nobel Prize and established light quanta as fundamental carriers of electromagnetic energy. Further evidence for photons as particles came from Compton scattering, observed in 1923 when X-rays scattered off electrons in light elements showed a wavelength shift dependent on scattering angle. explained this as an between a of h\nu and h\nu/c and a , analogous to billiard balls, yielding a wavelength change \Delta\lambda = \frac{h}{m_e c} (1 - \cos\theta), where m_e is the , c is the , and \theta is the scattering angle. This shift, unexplainable by classical wave scattering, confirmed the corpuscular of light quanta and earned the 1927 . The quantum nature of light exemplifies wave-particle duality, where light behaves as both waves and particles depending on the experimental context. Extending this duality to matter, Louis de Broglie hypothesized in 1924 that particles like electrons possess wave properties with wavelength \lambda = h/p, where p is momentum; this relation, initially proposed for photons (p = h/\lambda), provided a symmetric framework for duality and was experimentally verified through electron diffraction. De Broglie's insight laid the groundwork for wave mechanics, illustrating how light's dual nature mirrors that of matter particles. In , the imposes fundamental limits on measurements, capturing the interplay between wave and particle aspects. Formulated by in 1927, it states that the product of uncertainties in position \Delta x and wave number \Delta k (related to by p = \hbar k) satisfies \Delta x \Delta k \geq 1/2, preventing simultaneous precise knowledge of a photon's position and direction. This principle explains resolution limits in optical imaging, such as the diffraction barrier, and underscores why quantum measurements inherently disturb the system, distinguishing from classical descriptions.

Lasers and Coherent Sources

Lasers represent a cornerstone of modern optics, enabling the generation of highly coherent light through the process of . In 1917, introduced the theoretical framework for this phenomenon by deriving the relationships between , , and using that quantify the probabilities of these atomic transitions. The Einstein A describes the rate of , where an excited atom randomly emits a without external influence, while the Einstein B govern both stimulated absorption (-induced from a lower state) and stimulated emission (-induced de-excitation from an upper state, producing an identical ). These reveal that under thermal equilibrium, the number of atoms in higher states is exponentially lower than in lower states, following the . To achieve net of light, a requires , a non-equilibrium condition where more atoms or molecules occupy a higher than a lower one, making dominate over . This inversion is maintained by an external source, such as or electrical discharge, which excites atoms to upper levels faster than they decay. In the absence of inversion, and would prevent coherent amplification. The LASER stands for Light by of Radiation, a concept first proposed theoretically by Arthur Schawlow and Charles Townes in 1958 for optical frequencies, building on earlier developments. Laser operation typically involves a gain medium, an optical resonator, and a pumping mechanism, with energy level schemes classified as three-level or four-level systems. In a three-level system, like the ruby laser, pumping raises electrons from the ground state to a high-energy band, from which they non-radiatively decay to a metastable upper laser level; lasing occurs as these decay back to the ground state, but achieving inversion requires exciting over half the population, making it inefficient. Four-level systems, such as the Nd:YAG laser, offer greater efficiency: pumping populates a higher level that decays to a metastable upper laser level, while the lower laser level lies above the ground state and quickly empties via thermal relaxation, allowing easier inversion with fewer excited atoms. This distinction enables continuous-wave operation in many four-level lasers. Common laser types vary by gain medium and application. Gas lasers, exemplified by the helium-neon (He-Ne) , use a low-pressure gas mixture excited by an ; the He-Ne produces visible at 632.8 nm and is valued for its stability and narrow linewidth, often used in alignment and . Solid-state lasers, such as the neodymium-doped yttrium aluminum garnet (Nd:YAG), employ a crystalline host doped with rare-earth ions, pumped by flashlamps or s to emit at 1064 nm in the near-; these are versatile for high-power operations due to their robustness and ability to produce short pulses. diode lasers, or laser diodes, operate via current injection across a p-n junction in materials like , achieving direct electrical pumping and compact sizes; they emit across a broad spectrum from to and dominate in and fiber optics for their efficiency and tunability. A defining feature of lasers is their high temporal coherence, quantified by the coherence length l_c = \frac{\lambda^2}{\Delta \lambda}, where \lambda is the central wavelength and \Delta \lambda is the spectral linewidth. This length indicates the maximum path difference over which the light wave maintains a fixed phase relationship, enabling precise applications like interferometry, where even small displacements can be measured with sub-wavelength accuracy. For instance, a He-Ne laser with \Delta \lambda \approx 1 GHz yields a coherence length of tens of meters, far exceeding that of incoherent sources. Stimulated emission, rooted in quantum principles, underpins this coherence by producing photons in phase and direction.

Nonlinear and Advanced Phenomena

Nonlinear optics encompasses phenomena where the optical response of a depends on the of the , arising when the is no longer linearly proportional to the . This emerged following the invention of lasers, which provide the high intensities necessary to observe such effects, typically on the order of megawatts per square centimeter or higher. The theoretical foundation rests on expanding the \mathbf{P} in a of the \mathbf{E}: \mathbf{P} = \epsilon_0 [\chi^{(1)} \mathbf{E} + \chi^{(2)} \mathbf{E}^2 + \chi^{(3)} \mathbf{E}^3 + \cdots], where \chi^{(n)} are the nth-order susceptibilities, with \chi^{(1)} governing linear optics and higher orders enabling nonlinear interactions. Second-order nonlinearity, characterized by the nonzero \chi^{(2)} tensor, occurs in noncentrosymmetric materials and permits processes like (SHG), where two photons of frequency \omega_1 combine to produce one at \omega_2 = 2\omega_1. SHG was first experimentally demonstrated in using a focused into a , producing at half the of the input beam, confirming the nonlinear response predicted by . In SHG, phase matching is crucial for efficient conversion, often achieved via or quasi-phase matching techniques, enabling applications in frequency doubling for blue-violet lasers. Third-order nonlinearity, described by \chi^{(3)}, is ubiquitous in all materials and leads to effects such as the optical , where the refractive index n varies with light intensity I: n = n_0 + n_2 I, with n_2 the nonlinear index coefficient related to \chi^{(3)} by n_2 = \frac{3}{4n_0 \epsilon_0 c} \Re[\chi^{(3)}]. This intensity-dependent index was first observed in 1964 in liquids like , using pulsed laser beams to induce measurable via changes. (SPM) arises from the when a propagates in a nonlinear medium, causing its to vary across the temporal profile due to the intensity gradient, broadening the spectrum and enabling applications in optical switching and supercontinuum generation. In optical fibers, the interplay between the Kerr nonlinearity and allows for the formation of , stable pulse shapes that maintain their form over long distances. These fundamental balance the self-focusing tendency of the against dispersive broadening, governed by the i \frac{\partial u}{\partial z} - \frac{1}{2} \beta_2 \frac{\partial^2 u}{\partial t^2} + \gamma |u|^2 u = 0, where \beta_2 is the dispersion parameter and \gamma the nonlinearity coefficient. The concept was theoretically proposed in 1973 for anomalous dispersion regimes in low-loss fibers, predicting stationary pulse propagation. Experimental observation followed in 1980 using color-center lasers, demonstrating transmission over hundreds of kilometers without distortion, revolutionizing high-bit-rate optical communications. Advanced phenomena extend to engineered materials like metamaterials, which exhibit effective refractive indices n < 0 through subwavelength structuring, leading to where light bends oppositely to . The theoretical possibility of negative n was explored in 1968, showing that simultaneous negative \epsilon < 0 and permeability \mu < 0 would reverse both and group velocities, enabling superlenses and . Experimental realization came in with a composite of split-ring resonators and wires at frequencies, achieving n = -2.70 over a passband, verified by refraction and transmission measurements. Plasmonics involves surface plasmons, collective electron oscillations at metal-dielectric interfaces, enabling subwavelength light confinement beyond the diffraction limit. These were first theoretically described in for thin metal films, predicting resonant excitations that couple light to waves, with k_{sp} = k_0 \sqrt{\frac{\epsilon_m \epsilon_d}{\epsilon_m + \epsilon_d}} for propagation constant k_{sp}. polaritons (SPPs) at interfaces, like gold-air, exhibit strong field enhancement, underpinning nanophotonic devices such as sensors and waveguides.

Applications of Optics

Optical Imaging and Instrumentation

Optical imaging systems harness the principles of light refraction and focusing to form detailed images of objects, ranging from microscopic structures to distant celestial bodies. These instruments, including the and engineered devices like microscopes, telescopes, and cameras, rely on lenses or mirrors to collect and converge rays onto a detection surface, such as the or a . The effectiveness of these systems is determined by factors like , , and , which enable visualization beyond the unaided eye's capabilities. The serves as the quintessential , with its optimized for . enters through the , a transparent dome-shaped structure that provides about two-thirds of the eye's total refractive power by bending incoming rays. Behind the cornea lies the aqueous humor, , and crystalline , which together focus light onto the . The , suspended by zonular fibers and controlled by ciliary muscles, undergoes to adjust its curvature, shortening its for near objects (typically down to 25 cm) and relaxing for distant ones. This process allows the eye to maintain sharp focus across a range of distances, with an effective of approximately 17 mm. The eye's , defined by the diameter, yields an ranging from about f/2.1 in dim light to f/8.3 in bright conditions, influencing light intake and . Microscopes extend the eye's reach to the nanoscale, employing compound lens configurations to achieve high linear magnification of small specimens. A typical compound microscope consists of an objective lens close to the object, forming a real, enlarged intermediate image, which is then magnified further by an eyepiece acting as a simple magnifier. The total magnification M is the product of the objective's lateral magnification M_{\text{obj}} and the eyepiece's angular magnification M_{\text{eyepiece}}, given by M = M_{\text{obj}} \times M_{\text{eyepiece}}, where M_{\text{obj}} = -L / f_{\text{obj}} (with L as the tube length and f_{\text{obj}} the objective focal length) and M_{\text{eyepiece}} = 25 \, \text{cm} / f_{\text{eyepiece}} for relaxed viewing. Common setups yield magnifications from 100× to 1000× or more. However, resolution—the ability to distinguish fine details—is limited by diffraction, with the minimum resolvable distance d approximated by the Rayleigh criterion as d = 0.61 \lambda / \text{NA}, where \lambda is the wavelength and NA is the numerical aperture of the objective (typically 0.1–1.4 for visible light, limiting resolution to about 0.2–0.5 μm). Higher NA objectives, often using immersion oils, push this limit but cannot exceed the fundamental diffraction barrier. Telescopes, in contrast, magnify angular size for viewing remote objects, using either refracting or reflecting designs to collect faint light over large apertures. Refracting telescopes employ an objective lens to form a real image at its focal plane, viewed through an eyepiece that produces a virtual image at infinity for relaxed observation; the angular magnification is M = -f_{\text{obj}} / f_{\text{eyepiece}}, where f_{\text{obj}} and f_{\text{eyepiece}} are the respective focal lengths, often yielding 10× to 500× or higher depending on the configuration. These systems suffer from chromatic aberration, mitigated by achromatic doublets. Reflecting telescopes avoid this issue by using curved mirrors as the primary optic, with common types including the Newtonian (parabolic primary mirror with flat secondary for side viewing), Cassegrain (concave primary and convex secondary for compact rear focus), and Ritchey-Chrétien (hyperbolic mirrors for reduced coma in large instruments). Reflectors dominate modern astronomy due to their scalability and ability to gather more light without dispersion. Cameras represent versatile artificial imaging systems, evolving from simple pinhole designs to sophisticated digital variants. The pinhole camera operates on the principle of rectilinear propagation of light through a small aperture (ideally ~0.1–1 mm), projecting an inverted, undistorted image onto a screen or film without lenses; its infinite depth of field arises from the geometric sharpness, though exposure times are long due to limited light throughput. Modern cameras incorporate lenses to increase light collection and control focus, with digital sensors—such as charge-coupled devices (CCDs) or complementary metal-oxide-semiconductor (CMOS) arrays—replacing film to capture images as pixelated charge distributions. Depth of field, the range of distances appearing acceptably sharp, is governed by the lens aperture (f-number), focal length, and sensor size; smaller apertures (higher f-numbers) extend it, as the circle of confusion on the sensor remains below the resolution threshold (typically ~2–4 pixels). For instance, a 50 mm lens at f/8 on a full-frame sensor provides a depth of field of several meters at typical subject distances, enabling applications from portraiture to landscape photography.

Communication and Information Processing

Optical fibers serve as the backbone of modern systems, guiding signals through at the core-cladding interface, where the of the core is higher than that of the surrounding cladding, confining the within the core for efficient long-distance transmission. In silica-based single-mode fibers, which are widely used in , typical is approximately 0.15–0.2 dB/km at the 1550 nm , primarily due to and material absorption, enabling signals to travel thousands of kilometers with minimal loss. Additionally, chromatic dispersion in these silica fibers, arising from the wavelength-dependent , is about 17 ps/(nm·km) at 1550 nm, which can broaden optical pulses over distance but is managed through dispersion-compensating techniques to maintain signal integrity. Wavelength-division multiplexing (WDM) enhances the capacity of optical fibers by simultaneously transmitting multiple independent data channels at distinct wavelengths within the same fiber, effectively multiplying the without requiring additional fibers. In dense WDM systems, channel spacings as narrow as 0.8 nm (100 GHz) allow for dozens to hundreds of channels in the C-band (1530–1565 nm), supporting aggregate data rates exceeding terabits per second over transoceanic distances when combined with erbium-doped fiber amplifiers. Photonic integrated circuits (PICs) enable compact, high-speed manipulation of optical signals on a chip-scale , integrating components such as waveguides and modulators to process directly in the optical , reducing and power consumption compared to electronic counterparts. Waveguides in PICs, often fabricated from or , confine and route via similar to fibers but on micrometer scales, while electro-optic modulators, such as Mach-Zehnder interferometers, encode onto by or shifts at speeds up to 100 Gb/s per channel through carrier depletion or effects. Holography contributes to optical information processing by recording and reconstructing data through patterns in volume media, offering high-density capacities far beyond traditional methods. In volume holograms, the between a reference and the signal —carrying multiplexed data pages—creates a three-dimensional modulation within photosensitive materials like photorefractive or polymers, allowing thousands of pages to be stored in a single volume via angle or . occurs by illuminating the hologram with the reference , diffracting to retrieve the original signal with minimal , enabling areal densities up to 515 Gb/in² in demonstrations and potential for parallel applications.

Scientific and Industrial Uses

is a cornerstone of scientific research in , enabling the analysis of material properties through the interaction of with matter. measures the attenuation of as it passes through a sample, revealing molecular structures based on wavelength-specific lines corresponding to or vibrational transitions. Emission spectroscopy, conversely, detects emitted by excited atoms or molecules, providing insights into energy levels and compositions, as seen in atomic emission spectra used for . These techniques are fundamental for identifying chemical bonds and studying quantum phenomena in gases, liquids, and solids. Fourier Transform Infrared (FTIR) spectroscopy extends these principles into the range, where it records spectra to characterize vibrational modes of molecules, aiding in the of functional groups in compounds. FTIR achieves high by interferometrically modulating and applying transforms to the resulting interferogram, offering rapid, non-destructive analysis essential for pharmaceutical and material science applications. spectroscopy complements FTIR by probing inelastic scattering, where incident photons exchange energy with molecular vibrations, producing shifted wavelengths that reveal symmetric vibrational modes inaccessible to . This technique, enhanced by lasers for signal amplification, is particularly valuable for analysis of aqueous samples and solids without water interference. Optical sensors leverage interferometric principles for precise in scientific and settings. Fabry-Perot interferometers consist of two reflective surfaces forming a resonant cavity, where changes in cavity length due to or alter the pattern of transmitted or reflected , enabling sub-micrometer detection. These sensors are widely used in for and in harsh environments owing to their compact, robust and immunity to . Fiber Bragg gratings (FBGs) inscribed in optical fibers reflect specific wavelengths determined by the period, which shifts with applied or via photoelastic and effects, respectively. FBGs facilitate multiplexed sensing networks for in bridges and pipelines, with typical sensitivities around 1 pm/µε and sensitivities of 10 pm/°C. In semiconductor manufacturing, optical lithography employs ultraviolet (UV) light to pattern microcircuits onto wafers, projecting mask features through reduction optics onto photoresist-coated substrates. Deep UV (DUV) exposure at 193 nm wavelengths, using lasers, achieves feature sizes down to 10 nm by optimizing and techniques, though diffraction limits resolution according to the Rayleigh criterion, approximately λ/(2NA), where λ is the and NA the . Challenges include managing line-edge roughness and overlay precision as scaling pushes beyond 3 nm nodes, necessitating to compensate for optical aberrations. Medical optics integrates these principles for diagnostic and therapeutic applications. utilizes flexible fiber-optic bundles or digital imagers to deliver and collect inside the body, enabling real-time visualization of internal organs with illumination and magnification for minimally invasive procedures like gastrointestinal examinations. , exemplified by (laser-assisted in situ keratomileusis), reshapes the using excimer lasers at 193 nm to correct refractive errors, creating a precise stromal profile that improves without incisions, achieving over 95% patient satisfaction in suitable candidates. () employs low-coherence to generate micrometer-resolution cross-sectional images of tissue, particularly in for retinal layer assessment and in for intravascular plaque characterization, with axial resolutions down to 1-15 µm depending on source . Additionally, , originally developed for astronomy to correct atmospheric via deformable mirrors and sensing, enhances high-resolution retinal imaging in medical contexts by compensating for ocular aberrations.

References

  1. [1]
    What is Optics? : About Us
    Optics is the study of light, how it is generated, propagated, and detected, and how it interacts with matter.
  2. [2]
    [PDF] 25 GEOMETRIC OPTICS
    In particular, optics is concerned with the generation and propagation of light and its interaction with matter.
  3. [3]
    [PDF] Physics of Light and Optics
    Topics are addressed from a physics perspective and include the propagation of light in matter, reflection and transmission at boundaries, polarization effects,.
  4. [4]
    [PDF] Geometric Optics - userhome.brooklyn...
    Optics is the branch of physics that deals with the behavior of visible light and other ... Ray. The word “ray” comes from mathematics and here means a straight ...
  5. [5]
    General Terms - Exploring the Science of Light
    Physical optics – The branch of optics which studies interference, diffraction, polarization and other phenomena for which the ray approximation of ...
  6. [6]
    Optics and Photonics: Essential Technologies for Our Nation (2013)
    Optics and photonics are central to modern life, enabling integrated circuits, displays, internet, and medical tools, with potential for future societal impact.
  7. [7]
    What is Optics & Photonics? - CREOL #ucf
    Photonics enables advanced guidance systems, surveillance, infrared visioning, and directed-energy weapons for military and defense applications. Research. The ...
  8. [8]
    History of Optics - UF Physics
    Brief Overview. The ancient Greeks and Arabs had some knowledge of. the nature and properties of light. · The Greeks. In optics, Euclid's textbook (called the ...
  9. [9]
    [PDF] brief history of optics; absorption, refraction - MIT OpenCourseWare
    Mar 9, 2020 · Cover the fundamental properties of light propagation and interaction with matter under the approximations of geometrical optics and scalar.
  10. [10]
    Anatomy of an Electromagnetic Wave - NASA Science
    Aug 3, 2023 · Light is made of discrete packets of energy called photons. Photons carry momentum, have no mass, and travel at the speed of light. All light ...
  11. [11]
    Properties of Light - Tulane University
    Oct 17, 2014 · Light is electromagnetic radiation that has properties of waves. The electromagnetic spectrum can be divided into several bands based on the wavelength.
  12. [12]
    Wave-Particle Duality - HyperPhysics
    Most commonly observed phenomena with light can be explained by waves. But the photoelectric effect suggested a particle nature for light.
  13. [13]
    Light and Color - Electromagnetic Radiation - Molecular Expressions
    May 17, 2016 · Every category of electromagnetic radiation, including visible light, oscillates in a periodic fashion with peaks and valleys (or troughs), and ...
  14. [14]
    speed of light in vacuum - CODATA Value
    speed of light in vacuum $c$ ; Numerical value, 299 792 458 m s ; Standard uncertainty, (exact).
  15. [15]
    Kilogram: Mass and Planck's Constant | NIST
    May 14, 2018 · ... E = hν, the first "quantum" expression in history, stated by Max Planck in 1900. Here, E is energy, ν is frequency (the ν is not a “v” but ...
  16. [16]
    Visible Light - NASA Science
    Aug 4, 2023 · Violet has the shortest wavelength, at around 380 nanometers, and red has the longest wavelength, at around 700 nanometers.
  17. [17]
    Introduction - Richard Fitzpatrick
    The first law is the law of rectilinear propagation, which states that light rays propagating through a homogeneous transparent medium do so in straight-lines.
  18. [18]
    VIII. A dynamical theory of the electromagnetic field
    Oct 27, 2025 · A Dynamical Theory of the Electromagnetic Field. By J. Clerk Maxwell, F.B.S. ... PROFESSOR CLERK MAXWELL ON THE ELECTROMAGNETIC FIELD.
  19. [19]
    Heinrich Hertz - Magnet Academy - National MagLab
    In 1886, Hertz began experimenting with sparks emitted across a gap in a short metal loop attached to an induction coil. He soon built a similar apparatus, but ...
  20. [20]
    Maxwell's equations | Institute of Physics
    In 1865 Maxwell wrote down an equation to describe these electromagnetic waves. The equation showed that different wavelengths of light appear to us as ...
  21. [21]
    [PDF] Chapter 13 Maxwell's Equations and Electromagnetic Waves - MIT
    Thus, we conclude that light is an electromagnetic wave. The spectrum of electromagnetic waves is shown in Figure 13.4.4. Figure 13.4.4 Electromagnetic spectrum.
  22. [22]
    [PDF] Maxwell's Equations and EM Waves - UF Physics Department
    Nov 21, 2006 · Derivation of Electromagnetic Wave Equation. Now let's see how we can combine the differential forms of Maxwell's equations to derive a set ...
  23. [23]
    Classification of Polarization - HyperPhysics
    A plane electromagnetic wave is said to be linearly polarized. The transverse electric field wave is accompanied by a magnetic field wave as illustrated.
  24. [24]
  25. [25]
    Electromagnetic Waves and Polarization - NASA SVS
    Jul 7, 2017 · Circular polarization can be classified as left-circularly polarized or right-circularly polarized. There is also elliptical polarization which ...
  26. [26]
    The electromagnetic spectrum (article) | Khan Academy
    The electromagnetic spectrum is the range of EM radiation's wavelength, frequency, and energy, including radio, microwaves, infrared, visible light, ...
  27. [27]
    The Electromagnetic Spectrum: Exploring the Colors of Light
    The electromagnetic spectrum includes radio waves, microwaves, infrared, visible light, ultraviolet, X-rays, and gamma rays, each with different wavelengths ...
  28. [28]
    History of Geometric Optics
    Early ideas included light traveling from the eye, and infinite speed. Alhazan realized light travels to the eye. Snell discovered the true refraction law, and ...
  29. [29]
    Light through the ages: Ancient Greece to Maxwell - MacTutor
    Long before Lucretius, Euclid had made a mathematical study of light. He wrote Optica Ⓣ in about 300 BC in which he studied the properties of light which he ...
  30. [30]
    On Euclid and the Genealogy of Proof - Michigan Publishing
    Dec 13, 2021 · Abstract. I argue for an interpretation of Euclid's postulates as principles grounding the science of measurement.Missing: sources | Show results with:sources
  31. [31]
    Atmospheric refraction: a history - Optica Publishing Group
    The concept that the atmosphere could refract light entered Western science in the second century B.C. Ptolemy, 300 years later, produced the first clearly ...
  32. [32]
    A Very Brief History of Light | SpringerLink
    Dec 14, 2016 · Ptolemy carried out careful experiments on refraction and concluded that, for light propagating from one medium to another, the ratio of the ...
  33. [33]
    [PDF] Claudius Ptolemy (ca. AD 100 - arXiv
    Pseudo Euclidean reflections in ancient optics: A reexamination of textual issues pertaining to the Euclidean Optica and Catoptrica. Physis, 31, 1–. 45 ...
  34. [34]
    Ibn al-Haytham: 1000 Years after the Kitāb al-Manāẓir
    Oct 1, 2015 · Ibn al-Haytham attempted to demonstrate the intromission theory by showing how rays of light from an object proceed toward the eye along a ...
  35. [35]
    [PDF] THE SCIENCE OF OPTICS TO THE TIME OF KEPLER
    Ptolemy opens book 5 of the Optics by distinguishing reflection from refraction according to how the visual ray is broken in each case.
  36. [36]
    [PDF] Ibn Sahl's, Al- Haytham's and Young's works on refraction as ... - SPIE
    Aug 1, 2007 · One typical example of the impressive optical systems is the camera obscura invented by Ibn Al- Haytham. For techniques enabling us to easily ...
  37. [37]
    Optics to the Time of Kepler - Encyclopedia of the History of Science
    By contrast, the theory championed by Witelo and his perspectivist compeers has the crystalline lens visually sense the rays of light that impinge directly upon ...
  38. [38]
    Roger Bacon's Optics - AIP.ORG
    May 20, 2024 · The books featured were versions of 13th-century philosopher and scientist Roger Bacon 's seminal work Opus Majus, originally written in 1267 for Pope Clement ...
  39. [39]
    Roger Bacon - Stanford Encyclopedia of Philosophy
    Apr 19, 2007 · Roger Bacon (1214/1220–1292), Master of Arts, contemporary of Robert Kilwardby, Peter of Spain, and Albert the Great at the University of Paris in the 1240s.
  40. [40]
    Early Prints Depicting Eyeglasses - JAMA Network
    Spectacles had been invented in Pisa, Italy, around 1285. The first known illustration of them occurs in a mural in the chapter house of San Nicolò in Treviso, ...
  41. [41]
    Studying the Foundations of Optics with the Master - AIP Publishing
    Dec 1, 2023 · Perhaps facilitated by obscuring the Arab origins of the theory, Witelo's and Pecham's works became standard medieval European university texts.
  42. [42]
    The First Telescopes (Cosmology - American Institute of Physics
    A variation on the Galilean telescope was suggested by Johannes Kepler in his 1611 book Dioptrice. He noted that a telescopic device could be built using two ...
  43. [43]
    Johannes Kepler - The Galileo Project | Science
    Kepler went on to provide the beginning of a theory of the telescope in his Dioptrice, published in 1611. During this period the Keplers had three children ...
  44. [44]
    [PDF] Descartes' Dioptrics | Scholars at Harvard
    In the second discourse, Descartes attempts to derive the law of reflection (known since antiquity) and the law of refraction (first published in the Dioptrics ...
  45. [45]
    Descartes, Newton, and Snell's law - Optica Publishing Group
    IV. DESCARTES' DERIVATION OF THE PSEUDO-SNELL LAW. When Descartes published La Dioptrique in 1637, light refraction was a well-known and fairly accurately ...
  46. [46]
    The Project Gutenberg eBook of Treatise on Light, by Christiaan ...
    Treatise on Light in which are explained the causes of that which occurs in reflexion, & in refraction and particularly in the strange refraction of Iceland ...Missing: primary | Show results with:primary
  47. [47]
    [PDF] Christiaan Huygens' Wave Theory of Light: A Major Contribution to ...
    The world- renowned Dutch scientist from Den Haag, Christiaan. Huygens, developed the wave theory of light as depicted in his 1690 publication of Treatise On ...Missing: primary | Show results with:primary
  48. [48]
    [PDF] Isaac Newton (1642-1727) - PhilArchive
    Robert Hooke objected that Newton's account improperly assumed a corpuscular theory of light. Hooke held that Newton's account of light could not be certain ...
  49. [49]
    Newton's optics and atomism (Chapter 7)
    Jul 5, 2016 · While it is true that Newton believed in a corpuscular theory, utilized it in developing many of his optical experiments and theories, and ...Missing: Isaac | Show results with:Isaac
  50. [50]
    Thomas Young and the Nature of Light - American Physical Society
    He did do some further work in physics, and in 1807, Young published some of his lectures, including the double- slit version of the interference experiment.
  51. [51]
    27.3: Young's Double Slit Experiment - Physics LibreTexts
    Feb 20, 2022 · Young's double slit experiment gave definitive proof of the wave character of light. An interference pattern is obtained by the ...
  52. [52]
    July 1816: Fresnel's Evidence for the Wave Theory of Light
    But by the early 19th century, the wave theory was making a comeback, thanks in part to the work of a French civil engineer named Augustin-Jean Fresnel. Born in ...
  53. [53]
    First- and second-order Poisson spots | American Journal of Physics
    Aug 1, 2009 · In 1818 Fresnel submitted a treatise, Memoir on the Diffraction of Light,1 to the French Academy claiming that first-order diffraction effects ...
  54. [54]
    July 1849: Fizeau Publishes Results of Speed of Light Experiment
    Jul 1, 2010 · He knew the accepted value for the diameter of Earth's orbit at that time, and from that, he concluded that the speed of light was 240,000 ...
  55. [55]
    Speed of Light - SPIE
    Figure 2.3 The first terrestrial measurement of the speed of light was done by Fizeau in 1849 when he projected a pulsed beam of light onto a distant mirror.
  56. [56]
    speed of light - Science, civilization and society
    Fizeau performed the experiment between Suresnes and Montmartre in Paris over a distance of 8,633 m and obtained a value of 315,300 km/second, within about 4 ...
  57. [57]
    Léon Foucault - The Society of Catholic Scientists
    Foucault and Fizeau, working independently, were the first to make accurate direct measurements of the speed of light. ... Foucault measured it in 1850, and by ...
  58. [58]
    VIII. A dynamical theory of the electromagnetic field - Journals
    Maxwell James Clerk. 1865VIII. A dynamical theory of the electromagnetic fieldPhil. Trans. R. Soc.155459–512http://doi.org/10.1098/rstl.1865.0008. Section.
  59. [59]
    A dynamical theory of the electromagnetic field - Smithsonian Libraries
    Maxwell published this, his first classic paper on the elctromagnetic field in 1865. He developed the concept of electromagnetic radiation.
  60. [60]
    Milestones:Maxwell's Equations, 1860-1871
    James Clerk Maxwell conceived and developed his unified theory of electricity, magnetism and light. A cornerstone of classical physics.
  61. [61]
    26 Optics: The Principle of Least Time - Feynman Lectures - Caltech
    The simplest object is a mirror, and the law for a mirror is that when the light hits the mirror, it does not continue in a straight line, but bounces off the ...
  62. [62]
    3.Fermat's Principle of Least Time - Galileo and Einstein
    The ray that enters your eye from the mirror goes along the shortest bouncing-off-the-mirror path. You can prove that this is equivalent to angle of incidence ...
  63. [63]
    [PDF] Fermat's Principle and the Laws of Reflection and Refraction ( )2
    The derivations are given below. (a) Consider the light ray shown in the figure. A ray of light starting at point A reflects off the surface at point ...
  64. [64]
    Total Internal Reflection - HyperPhysics Concepts
    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: source | Show results with:source
  65. [65]
    Refraction - A Green Flash Page
    The actual law of refraction was discovered in the early 1600s by a Dutch mathematician and geodesist, Willebrord Snel van Royen. (Because his name in Latin is ...Missing: derivation | Show results with:derivation
  66. [66]
    Thin-Lens Equation:Cartesian Convention - HyperPhysics
    The derivation of the Gaussian form proceeds from triangle geometry. For a thin lens, the lens power P is the sum of the surface powers.
  67. [67]
    Mirror Equation - HyperPhysics
    Spherical Mirror Equation. The equation for image formation by rays near the optic axis (paraxial rays) of a mirror has the same form as the thin lens equation ...Missing: historical source
  68. [68]
  69. [69]
    pass cells based on a spherical mirror aberration
    rays make small angles to the optic axial of the system, so that three important angle approximations are valid, i.e., sin θ ≈ θ, tan θ ≈ θ, and cos θ ≈ 1.
  70. [70]
    [PDF] Paraxial focal length measurement method with a simple apparatus
    Jan 23, 2019 · Once the angle between incident ray and optical axis is smaller than 15 degrees, the relative error of paraxial approximation is below 1%. If ...
  71. [71]
    [PDF] Paraxial Ray Optics Cloaking - Chapman University Digital Commons
    Nov 18, 2014 · The paraxial equations we presented are valid for angles θ that satisfy tanθ ≈ sinθ ≈ θ. This is true for θ = 10◦ −15◦, while even θ up ...
  72. [72]
    [PDF] Tutorial: Geometrical Optics and Ray Tracing - UR Research
    May 2, 2011 · They're derived from the basic geometry (say radii of curvature, distance) and basic physics (Snell's law, indices of refraction). The results.
  73. [73]
    [PDF] Chapter on - Geometrical Optics - Physics2000
    Lens design is an ideal problem for the computer, for tracing light rays through a lens system requires many repeated applica- tions of Snell's law. When we ...
  74. [74]
    Chapter 2 - Skew-Ray Tracing of Geometrical Optics
    Geometrical optics in this book are applied to perform sequential raytracing based on Snell's law and a homogeneous coordinate notation, and then compute ...<|control11|><|separator|>
  75. [75]
    [PDF] 2.5 Rays and Optical Systems
    More complicated imaging systems, such as thick lenses, can be described by ray matrices and arbitrary paraxial optical systems can be analyzed with them ...
  76. [76]
    ABCD Matrix Analysis Tutorial/Ray Transfer Matrix ... - BYU Photonics
    An ABCD matrix is a 2x2 matrix that describes an optical element's effect on a light ray, allowing the output to be written in terms of its input.
  77. [77]
    [PDF] The ABCD matrices for reflection and refraction for any incident ...
    Oct 22, 2021 · ABCD matrices are 4x4 matrices for reflection and refraction at any incident angle on any surface, derived for any diopter.
  78. [78]
    [PDF] Gaussian Beam Optics - Experimentation Lab
    Self shows a method to model transformations of a laser beam through simple optics, under paraxial conditions, by calculating the Rayleigh range and beam waist ...
  79. [79]
    [PDF] Introduction to Gaussian Beams
    This can be accomplished by an approximate solution of the wave equation under what is known as the “paraxial approximation” and the corresponding “paraxial ...
  80. [80]
    [PDF] 3.0 Wave Equation
    are solutions to the wave equation and they represent spherical waves propagating outward from the origin. 3.5 Linear Superposition. Since the electric field ...
  81. [81]
    [PDF] Huygens principle; young interferometer; Fresnel diffraction
    sources; Huygens principle says that the transmitted. Huygens wavefront may also be decomposed into point sources. point sources. If the transparency is ...
  82. [82]
    Group Velocity - RP Photonics
    The group velocity of light in a medium is defined as the inverse of the derivative of the wavenumber with respect to angular optical frequency.
  83. [83]
    Malus's Law - SPIE
    In 1808, using a calcite crystal, Malus discovered that natural incident light became polarized when it wasreflected by a glass surface.Missing: seminal | Show results with:seminal
  84. [84]
    Etienne-Louis Malus: The Polarization of Light by Refraction and ...
    Aug 7, 2025 · This paper presents the memoir of Etienne-Louis Malus on a phenomenon linked to light polarization in birefringent materials and in ordinary reflection.Missing: seminal | Show results with:seminal
  85. [85]
  86. [86]
    Principles of Birefringence | Nikon's MicroscopyU
    In Figure 3, the incident light rays giving rise to the ordinary and extraordinary rays enter the crystal in a direction that is oblique with respect to the ...
  87. [87]
    [PDF] Birefringence.pdf
    When light rays are transmitted through a plate of anisotropic material as described above, the two rays refracted at the first surface refract again at the ...
  88. [88]
    Anomalous Dispersion - Ocean Optics Web Book
    Oct 14, 2021 · ... normal life at visible wavelengths is called “anomalous dispersion.” There is really nothing normal or abnormal about either situation, you ...Missing: prism authoritative
  89. [89]
    Opticks: or, A treatise of the reflections, refractions, inflexions and ...
    Jan 24, 2018 · A treatise of the reflections, refractions, inflexions and colours of light. Newton, Isaac. Printed for Sam. Smith, and Benj. Walford, London, 1704.
  90. [90]
    Blue Sky and Rayleigh Scattering - HyperPhysics
    This scattering, called Rayleigh scattering, is more effective at short wavelengths (the blue end of the visible spectrum). Therefore the light scattered down ...Missing: ∝ λ⁴ seminal paper
  91. [91]
    L. Rayleigh, “On the Light from the Sky, Its Polarization and Color ...
    L. Rayleigh, “On the Light from the Sky, Its Polarization and Color,” Philosophical Magazine, Vol. 41, 1871, pp. 107-120.
  92. [92]
    Mie Scattering - LaVision
    Mie scattering is elastic scattered light of particles that have a diameter similar to or larger than the wavelength of the incident light.Missing: λ⁴ seminal paper
  93. [93]
    [PDF] Einstein in 1916: “On the Quantum Theory of Radiation” - arXiv
    Mar 23, 2017 · 3The spontaneous emission has in the average no preferred direction, and therefore does in the average not transmit momentum to the atom. 3 ...
  94. [94]
    Three-Level and Four-Level Lasers - SPIE Digital Library
    In a three-level laser, a molecule is pumped from the ground state to an upper state. The upper state quickly decays by some nonradiative method.
  95. [95]
  96. [96]
    Coherence Length - RP Photonics
    The coherence length is a measure of temporal coherence, expressed as the propagation distance over which the coherence significantly decays.What is a Coherence Length? · Calculations · Lasers with Long Coherence...Missing: λ²/ Δλ
  97. [97]
    Generation of Optical Harmonics | Phys. Rev. Lett.
    Oct 31, 2014 · In 1961, researchers showed that laser light could be converted from one color to another, the first nonlinear optical effect.
  98. [98]
    Transmission of stationary nonlinear optical pulses in dispersive ...
    Research Article| August 01 1973. Transmission of stationary nonlinear optical pulses in ... A. Hasegawa and F. D. Tappert (unpublished). 8. V. J.. Karpman.
  99. [99]
    THE ELECTRODYNAMICS OF SUBSTANCES WITH ... - IOP Science
    THE ELECTRODYNAMICS OF SUBSTANCES WITH SIMULTANEOUSLY NEGATIVE VALUES OF AND μ. Viktor G Veselago. © 1968 American Institute of PhysicsMissing: refraction | Show results with:refraction
  100. [100]
    Plasma Losses by Fast Electrons in Thin Films
    Plasma Losses by Fast Electrons in Thin Films. R. H. Ritchie. Phys. Rev. 106, 874 – Published 1 June 1957. Article has an altmetric score of 3.Missing: plasmons | Show results with:plasmons
  101. [101]
    Refractive Telescopes - HyperPhysics
    Its length is equal to the sum of the focal lengths of the objective and eyepiece, and its angular magnification is -fo /fe , giving an inverted image. The ...
  102. [102]
    26.4 Microscopes – College Physics - UCF Pressbooks
    Calculate the magnification of an object placed 6.20 mm from a compound microscope ... Derived by mathematical formula. N A = n sin α ,. where n ...<|separator|>
  103. [103]
    Gross Anatomy of the Eye by Helga Kolb - Webvision
    Jan 25, 2012 · A transparent external surface, the cornea, that covers both the pupil and the iris. This is the first and most powerful lens of the optical ...
  104. [104]
    The Human Eye - HyperPhysics
    The inner lens is held in place by the ciliary fibers, which are important in the process of accommodation for close vision.
  105. [105]
    eye - OMLC
    The typical eye has a focal length of 17.3 mm, a variable f-number from f/2.5 to f/17, and adjustable focus from infinity down to 10 inches. While the full ...
  106. [106]
    [PDF] ASTR469 Lecture 10: Telescopes and Optics I (Ch. 6) - Sarah Spolaor
    Feb 6, 2019 · For instance, the f-number of the human eye varies from about f/8.3 in a very brightly lit place to about f/2.1 in the dark. 3.2 Resolving power.
  107. [107]
    Microscope - HyperPhysics
    A compound microscope uses a very short focal length objective lens to form a greatly enlarged image. This image is then viewed with a short focal length ...
  108. [108]
    Compound microscope - Total magnification - The Virtual Edge
    Objective lens X Ocular lens = Total magnification ; For example: low power: (10X)(10X) = 100X ; high dry: (40X)(10X) = 400X ; oil immersion: (100X)(10X) = 1000X ...
  109. [109]
    27.6 Limits of Resolution: The Rayleigh Criterion - UCF Pressbooks
    The Rayleigh criterion for the diffraction limit to resolution states that two images are just resolvable when the center of the diffraction pattern of one is ...
  110. [110]
    Microscopy Basics | Numerical Aperture and Resolution
    Resolution is directly related to the useful magnification of the microscope and the perception limit ... limit that is referred to as the Rayleigh criterion. A ...
  111. [111]
    telescope_keplerian.html - UNLV Physics
    The formula for telescope magnification is M = -f_p/f_e . Since magnification is a ratio of like quantities, it can be regarded as a dimensionless number: i.e., ...
  112. [112]
    Pre-lab #5 Telescopes - MTSU Physics
    Figure 3.5 Reflecting Telescopes Four reflecting telescope designs: (a) prime focus, (b) Newtonian focus, (c) Cassegrain focus, and (d) coudé focus. Each ...
  113. [113]
    [PDF] Telescopes come in three basic styles
    Telescopes come in three basic styles: refracting (using lenses), reflecting (using mirrors), and catadioptric (using mirrors and lenses).
  114. [114]
    [PDF] The Pinhole Camera and Image Resolution
    Now think about Huygens' principle, where each point inside the pinhole can be thought of as a source of secondary spherical wavelets. (see Figure 2). The waves ...
  115. [115]
    The Pinhole: Nature's Lens - Geophysical Institute
    May 18, 1986 · Pinhole cameras, despite their simplicity, provide nearly distortion-free images and an almost infinite depth of field.
  116. [116]
    Depth of field - Stanford Computer Graphics Laboratory
    Mar 1, 2012 · The standard definition is that the width of the blurred image of the feature has become larger on the sensor than some maximum allowable circle of confusion.
  117. [117]
    [PDF] Section 17 Camera Systems
    The separation between the sensor and the rear focal point is given by the Depth of Focus. Objects at the Near Point will focus a Depth of Focus behind the ...
  118. [118]
  119. [119]
  120. [120]
  121. [121]
  122. [122]
  123. [123]
    Volume holographic data storage at an areal density of 250 ...
    Apr 1, 2001 · An optical system for high-density holographic data storage achieves a small Nyquist aperture by use of large SLM pixels and short-focal-length ...
  124. [124]
    Holographic digital data storage in a photorefractive polymer
    We have demonstrated a variety of digital data page writing, reading, and erasing operations in our photorefractive polymer samples, using a precision test ...
  125. [125]
    Spectroscopy: A Measurement Powerhouse | NIST
    Mar 28, 2025 · In absorption spectroscopy, scientists measure light that has passed through clouds of atoms or molecules. Each type of atom or molecule absorbs ...Missing: fundamentals | Show results with:fundamentals
  126. [126]
    Spectroscopy 101 – How Absorption and Emission Spectra Work
    Let's go back to simple absorption and emission spectra. We can use a star's absorption spectrum to figure out what elements it is made of based on the colors ...Missing: fundamentals | Show results with:fundamentals
  127. [127]
    Fourier Transform Infrared Spectrometer - an overview
    FTIR, or Fourier Transform Infrared Spectroscopy, is defined as an analytic technique that utilizes infrared light to measure the absorption of specific ...
  128. [128]
    Raman Techniques: Fundamentals and Frontiers - PMC
    The Raman effect originates from the inelastic scattering of light, and it can directly probe vibration/rotational-vibration states in molecules and materials.
  129. [129]
    Chronology of Fabry-Perot Interferometer Fiber-Optic Sensors ... - NIH
    In this study, a wide variety of FPI sensors are reviewed in terms of fabrication methods, principle of operation and their sensing applications.
  130. [130]
    Fibre Bragg Grating Based Strain Sensors: Review of Technology ...
    Sep 15, 2018 · In this paper we review FBG strain sensors with high focus on the underlying physical principles, the interrogation, and the read-out techniques.
  131. [131]
    Advancements and challenges in inverse lithography technology
    Jul 24, 2025 · The optical lithography exposure has evolved through three main stages: contact, proximity, and projection. Early IC manufacturing relied on ...
  132. [132]
    Evolution in Lithography Techniques - PubMed Central - NIH
    Optical lithography can be used up to 100 nm integrated circuit (IC). For semiconductor devices < 100 nm, next-generation lithography is required. Extreme ...
  133. [133]
    Advances in optical gastrointestinal endoscopy: a technical review
    In this review, we focus on the technical aspects and clinical applicability of optical endoscopy, highlighting recent advances in device and optical design, ...
  134. [134]
    Laser In Situ Keratomileusis (LASIK) - StatPearls - NCBI Bookshelf
    LASIK surgery changes the refractive power of the cornea first by creating a hinged corneal flap from the epithelium, Bowman's membrane, and the superficial ...
  135. [135]
    Optical Coherence Tomography: An Emerging Technology for ...
    Optical coherence tomography (OCT) is an emerging technology for performing high-resolution cross-sectional imaging. OCT is analogous to ultrasound imaging, ...