Electromagnetic spectrum
The electromagnetic spectrum is the complete range of electromagnetic radiation, consisting of waves formed by oscillating electric and magnetic fields that propagate through space at the constant speed of light in vacuum, defined as exactly 299,792,458 meters per second.[1][2] This spectrum spans a continuous distribution of frequencies and wavelengths, from extremely low-frequency radio waves (longer than 1 meter) with energies below 2 × 10^{-24} joules to gamma rays (shorter than 10^{-11} meters) with energies exceeding 2 × 10^{-14} joules, encompassing all forms of photon-based energy transfer.[3][4] The spectrum is conventionally divided into seven main regions based on wavelength, frequency, and typical applications or sources: radio waves, microwaves, infrared, visible light, ultraviolet, X-rays, and gamma rays.[4] Radio waves, with wavelengths greater than 0.1 meters and frequencies below 3 × 10^9 Hz, are used in communications and broadcasting.[3][5] Microwaves (wavelengths 1 × 10^{-3} to 1 × 10^{-1} meters) penetrate clouds and are essential for radar and satellite transmissions.[4] Infrared radiation (7 × 10^{-7} to 1 × 10^{-3} meters) is associated with thermal emission and detected as heat.[3] The narrow visible light band (4 × 10^{-7} to 7 × 10^{-7} meters) is the only portion perceivable by the human eye, spanning colors from violet to red.[6] Ultraviolet (UV) light (1 × 10^{-8} to 4 × 10^{-7} meters) includes ionizing wavelengths harmful to living cells, while X-rays (1 × 10^{-11} to 1 × 10^{-8} meters), produced by electron transitions in atoms, and gamma rays (below 1 × 10^{-11} meters), produced by nuclear processes, are high-energy, penetrating forms of radiation, often absorbed by Earth's atmosphere.[4][3][7] Key properties of electromagnetic waves include their transverse nature, ability to travel through vacuum without a medium, and inverse relationship between wavelength (λ) and frequency (f), governed by c = fλ, where c is the speed of light.[2] Photon energy increases with frequency (E = hf, with h as Planck's constant), making shorter-wavelength radiation more energetic and potentially ionizing, which has implications for biological effects, materials science, and astronomical observations.[3] Applications span telecommunications, medical imaging, remote sensing, and spectroscopy, with NASA's missions utilizing the full spectrum to explore the universe.[8]Fundamentals of Electromagnetic Waves
Nature and Propagation of Electromagnetic Waves
Electromagnetic waves are self-propagating transverse waves composed of oscillating electric and magnetic fields that are perpendicular to each other and to the direction of propagation./University_Physics_II_-Thermodynamics_Electricity_and_Magnetism(OpenStax)/16%3A_Electromagnetic_Waves/16.02%3A_Maxwells_Equations_and_Electromagnetic_Waves) The electric field \mathbf{E} and magnetic field \mathbf{B} vary sinusoidally in phase, with their mutual perpendicularity ensuring that the wave's energy transport occurs along the propagation axis without requiring a material medium.[2] This transverse nature distinguishes electromagnetic waves from longitudinal mechanical waves, such as sound, which rely on particle displacement parallel to the propagation direction and necessitate a medium for transmission./University_Physics_II_-Thermodynamics_Electricity_and_Magnetism(OpenStax)/16%3A_Electromagnetic_Waves/16.01%3A_The_Production_of_Electromagnetic_Waves) In vacuum, electromagnetic waves propagate at the constant speed of light, c = 299\,792\,458 m/s, which serves as a fundamental physical constant and the maximum speed for information transfer in the universe.[9] This velocity arises from the interplay of electric permittivity \epsilon_0 and magnetic permeability \mu_0 of free space, where c = 1 / \sqrt{\mu_0 \epsilon_0}./University_Physics_II_-Thermodynamics_Electricity_and_Magnetism(OpenStax)/16%3A_Electromagnetic_Waves/16.02%3A_Maxwells_Equations_and_Electromagnetic_Waves) Unlike mechanical waves, electromagnetic waves do not diminish in speed due to the absence of a medium, allowing them to traverse interstellar distances with minimal attenuation in empty space.[2] The theoretical foundation for electromagnetic waves is provided by Maxwell's equations, a set of four coupled differential equations that unify electricity, magnetism, and optics.[10] In their differential form for vacuum (where current density \mathbf{J} = 0), they are: \nabla \cdot \mathbf{E} = 0, \quad \nabla \cdot \mathbf{B} = 0, \nabla \times \mathbf{E} = -\frac{\partial \mathbf{B}}{\partial t}, \quad \nabla \times \mathbf{B} = \mu_0 \epsilon_0 \frac{\partial \mathbf{E}}{\partial t}. These equations predict wave solutions where changing electric fields induce magnetic fields and vice versa, enabling self-sustaining propagation.[11] Taking the curl of Faraday's law (\nabla \times \mathbf{E} = -\partial \mathbf{B}/\partial t) and substituting Ampère's law with Maxwell's correction yields the wave equation \nabla^2 \mathbf{E} = \mu_0 \epsilon_0 \partial^2 \mathbf{E}/\partial t^2, confirming the wavelike behavior./University_Physics_II_-Thermodynamics_Electricity_and_Magnetism(OpenStax)/16%3A_Electromagnetic_Waves/16.02%3A_Maxwells_Equations_and_Electromagnetic_Waves) When electromagnetic waves interact with matter, they exhibit behaviors such as reflection, refraction, diffraction, and polarization, governed by the material's dielectric properties and boundaries.[12] Reflection occurs at interfaces where the wave bounces off, with the angle of incidence equaling the angle of reflection, as described by Fresnel equations./06%3A_An_Introduction_to_Spectrophotometric_Methods/6.02%3A_Wave_Properties_of_Electromagnetic_Radiation) Refraction involves bending upon entering a medium with different refractive index n = c/v, where v is the wave speed in the material, following Snell's law n_1 \sin \theta_1 = n_2 \sin \theta_2.[13] Diffraction allows waves to bend around obstacles or through apertures, revealing their wave nature beyond geometric optics, while polarization refers to the orientation of the electric field vector, which can be linear, circular, or elliptical, and is altered by interactions with anisotropic media or polarizers.[12] These phenomena enable applications from optical instruments to wireless communication./University_Physics_II_-Thermodynamics_Electricity_and_Magnetism(OpenStax)/16%3A_Electromagnetic_Waves/16.01%3A_The_Production_of_Electromagnetic_Waves)Key Characteristics: Wavelength, Frequency, and Energy
The electromagnetic spectrum encompasses waves distinguished by two primary characteristics: wavelength and frequency. Wavelength, denoted as \lambda, represents the spatial distance between consecutive crests (or troughs) of the wave, typically measured in units ranging from meters for long radio waves to nanometers or angstroms ($1angstrom= 10^{-10}m) for shorter wavelengths in the ultraviolet, X-ray, and gamma-ray regions.[3] Frequency, denoted asf, quantifies the number of wave cycles occurring per second, expressed in hertz (Hz), with the spectrum spanning from low frequencies around 3 \times 10^4Hz for very low-frequency radio waves to frequencies exceeding3 \times 10^{19}$ Hz for gamma rays.[3]/Spectroscopy/Fundamentals_of_Spectroscopy/Electromagnetic_Radiation) These properties are fundamentally interrelated through the wave equation derived from Maxwell's theory of electromagnetism, which states that the speed of electromagnetic waves in vacuum, c, equals the product of frequency and wavelength: c = f \lambda.[14] Here, c is a universal constant with the exact value $299792458m/s.[9] This relation implies an inverse proportionality between [frequency](/page/Frequency) and [wavelength](/page/Wavelength): as [frequency](/page/Frequency) increases across the [spectrum](/page/Spectrum), [wavelength](/page/Wavelength) decreases proportionally, sincecremains fixed in [vacuum](/page/Vacuum). For instance, radio waves exhibit long [wavelengths](/page/Wavelength) (meters to kilometers) and low [frequencies](/page/Frequency) (kHz to MHz), while gamma rays have extremely short [wavelengths](/page/Wavelength) (less than10^{-12}m) and high [frequencies](/page/Frequency) (above10^{19}$ Hz).[3] A third key characteristic is the energy associated with electromagnetic radiation, particularly when viewed through the quantum lens where light behaves as discrete packets called photons. The energy E of a single photon is given by E = h f, where h is Planck's constant, with the exact value h = 6.62607015 \times 10^{-34} J s.[15] This formula, introduced by Einstein to explain the photoelectric effect, highlights the quantum nature of electromagnetic waves: higher-frequency photons carry more energy, enabling phenomena like ionization in the ultraviolet and gamma-ray regions, whereas lower-frequency photons, such as those in radio waves, have minimal individual energy but can accumulate through many photons. The electromagnetic spectrum thus forms a continuous progression ordered by increasing frequency (and energy) from radio waves to gamma rays, with no abrupt boundaries but rather a seamless transition governed by these interrelated properties.[16]Historical Development
Theoretical Foundations
The foundations of electromagnetic theory emerged in the early 19th century, beginning with Hans Christian Ørsted's 1820 discovery that an electric current flowing through a wire produces a magnetic field around it, thereby establishing a direct link between electricity and magnetism.[17] This observation, reported in Ørsted's pamphlet Experimenta circa effectum conflictus electrici in acum magneticam, demonstrated that the magnetic effect circled the current in a manner consistent with the right-hand rule, overturning the prior view of electricity and magnetism as separate phenomena.[18] Building on Ørsted's work, Michael Faraday conducted experiments in 1831 that revealed electromagnetic induction, showing that a changing magnetic field near a conductor induces an electric current within it.[19] Faraday's investigations, detailed in his paper "On the Induction of Electric Currents," involved moving magnets relative to coils or varying currents to produce induced electromotive forces, and he conceptualized these effects through the notion of continuous "lines of force" or fields permeating space, rather than relying solely on action-at-a-distance.[20] This field concept provided a qualitative framework for understanding how magnetic fields could influence electric charges without physical contact, paving the way for a unified theory. The culmination of these ideas came in James Clerk Maxwell's 1865 paper, "A Dynamical Theory of the Electromagnetic Field," which mathematically synthesized electricity, magnetism, and optics into a coherent framework.[21] Maxwell modified Ampère's circuital law by introducing the "displacement current" term, representing the rate of change of electric displacement in regions without conduction current, such as between capacitor plates; this addition ensured the continuity of current in circuits and enabled the propagation of electromagnetic disturbances as waves.[22] From his equations, Maxwell derived that these waves travel through space at a speed equal to that of light, approximately 3 × 10^8 meters per second in vacuum, leading him to conclude that light itself is an electromagnetic wave, thereby resolving the longstanding divide between optical phenomena and electrical-magnetic interactions.[23] This theoretical prediction was later experimentally verified by Heinrich Hertz in 1887 through the generation and detection of radio waves.[24]Experimental Discoveries
Prior to the unification of electromagnetic theory, experimental extensions of the visible spectrum had already revealed invisible regions. In 1800, British astronomer William Herschel discovered infrared radiation by passing sunlight through a prism and measuring temperatures with thermometers placed beyond the red end of the visible spectrum, finding higher heat in the invisible region, indicating longer-wavelength radiation associated with thermal energy.[25] In 1801, German physicist Johann Wilhelm Ritter identified ultraviolet radiation by observing that silver chloride paper blackened more rapidly beyond the violet end of the spectrum when exposed to sunlight dispersed by a prism, demonstrating shorter-wavelength, chemically active rays invisible to the eye.[26] These discoveries expanded the known optical spectrum and anticipated the broader electromagnetic nature of light. In 1887, Heinrich Hertz conducted pioneering experiments that experimentally verified the existence of electromagnetic waves predicted by James Clerk Maxwell's theory. Using a spark-gap transmitter consisting of an induction coil connected to a dipole antenna with a small gap, Hertz generated high-frequency oscillations, producing radio waves with wavelengths around 66 cm. He detected these waves with a simple loop receiver equipped with another spark gap, observing that the induced sparks confirmed the waves' propagation through space. Further tests demonstrated that these waves exhibited reflection from metal sheets and diffraction around obstacles, behaviors analogous to those of visible light, thus establishing radio waves as part of the electromagnetic spectrum.[27] Building on Hertz's findings, Oliver Lodge demonstrated practical transmission and detection of electromagnetic waves in 1894 during a lecture at the Royal Institution, using a coherer detector to receive signals over distances of about 150 meters and showcasing synthetic Hertzian waves for signaling purposes. In the mid-1890s, Guglielmo Marconi advanced this work by developing wireless telegraphy systems, filing a provisional patent in 1896 for a device that transmitted Morse code signals using spark transmitters and coherers, achieving ranges up to several kilometers by 1897 and extending operations to longer wavelengths for improved reliability over land and sea. These efforts marked the initial practical exploitation of radio waves, bridging laboratory demonstrations to communication applications.[28][29] In 1895, Wilhelm Röntgen discovered X-rays while investigating cathode rays in a vacuum tube, observing that when high-voltage electrons struck a glass wall or metal target, an unknown penetrating radiation emerged that could expose photographic plates and fluoresce screens even through opaque materials. He named these rays "X-rays" due to their unidentified nature and documented their ability to pass through soft tissues while being absorbed by denser structures like bones, laying the foundation for high-energy portions of the spectrum beyond ultraviolet. This serendipitous finding, reported in a preliminary communication to the Würzburg Physical-Medical Society, revolutionized imaging and spectral exploration.[30] Henri Becquerel accidentally discovered radioactivity in 1896 while studying phosphorescence in uranium salts, noticing that a uranium potassium sulfate crystal exposed a photographic plate even when shielded from light, indicating spontaneous emission of penetrating rays independent of excitation. Initially attributing this to a form of X-ray-like radiation from uranium, Becquerel's observations revealed continuous emission from the atomic nucleus through natural radioactive decay. His Comptes Rendus reports detailed the rays' ability to ionize air and discharge electroscopes. Subsequent studies identified components of this radiation, including gamma rays—highly energetic electromagnetic radiation with wavelengths shorter than X-rays—first observed by Paul Villard in 1900 from radium sources.[31] Philipp Lenard advanced understanding of the ultraviolet region in 1902 through meticulous photoelectric effect experiments, illuminating metal surfaces with UV light and measuring the ejected electrons' energies using retarding potentials. He found that electron emission occurred only above a threshold frequency specific to the metal, with kinetic energies increasing linearly with frequency rather than light intensity, challenging classical wave theory and suggesting a particle-like quantum interaction at UV wavelengths. These observations, conducted with improved vacuum tubes, provided early empirical hints at the quantized nature of electromagnetic energy in the near-UV spectrum.[32] In 1916, Robert Millikan quantitatively verified the quantum nature of light through photoelectric experiments, measuring stopping potentials for electrons ejected from clean metal surfaces under monochromatic UV and visible light to determine photon energies. Adapting techniques akin to his earlier oil-drop method for precision charge measurements, Millikan confirmed Einstein's equation E = h\nu - \phi, where h is Planck's constant, \nu is frequency, and \phi is the work function, yielding h = 6.57 \times 10^{-27} erg-second—close to modern values—and establishing discrete photon energies across the spectrum. His Physical Review paper provided rigorous data supporting the corpuscular model for electromagnetic waves in the optical and UV ranges.[33]Spectrum Classification
Criteria for Regional Division
The electromagnetic spectrum is conventionally divided into discrete regions using arbitrary yet practical boundaries defined by wavelength or frequency thresholds, which facilitate scientific study, technological applications, and standardization. These divisions emerged from a combination of physical properties, such as photon energy, and practical considerations including detection methods and historical conventions. For instance, the spectrum spans from radio waves with wavelengths longer than 1 mm (frequencies below 300 GHz) to gamma rays with wavelengths shorter than 0.01 nm (frequencies above 30 EHz), with intermediate regions like microwaves (1 mm to 1 m), infrared (700 nm to 1 mm), visible light (400–700 nm), ultraviolet (10–400 nm), and X-rays (0.01–10 nm).[34] Such boundaries are not rigid but reflect how electromagnetic waves interact differently with matter at various scales, influencing their propagation and detection—radio waves are easily detected by antennas, while X-rays require specialized detectors like scintillation counters./University_Physics_II_-Thermodynamics_Electricity_and_Magnetism(OpenStax)/16%3A_Electromagnetic_Waves/16.06%3A_The_Electromagnetic_Spectrum) A key criterion for regional division is the distinction between non-ionizing and ionizing radiation, based on photon energy thresholds that determine biological and chemical effects. Non-ionizing radiation, with energies below approximately 10–12 eV (corresponding to frequencies under 3 × 10¹⁵ Hz), includes radio waves, microwaves, infrared, visible light, and most ultraviolet; it lacks sufficient energy to eject electrons from atoms, instead causing thermal or vibrational effects. Ionizing radiation, with energies above this threshold (ultraviolet below ~100 nm, X-rays, and gamma rays), can ionize atoms by removing electrons, leading to potential cellular damage. This boundary falls within the ultraviolet region, where the transition is gradual rather than sharp, as ionization potential varies by material.[35][34] Atmospheric absorption significantly influences these divisions by creating transmission "windows" where electromagnetic waves propagate with minimal attenuation, guiding practical classifications and applications. Earth's atmosphere is largely transparent in the radio (5 MHz to 300 GHz), near- and mid-infrared (roughly 0.7–5 μm and 8–14 μm), and visible (0.4–0.7 μm) bands due to low absorption by gases like water vapor, oxygen, and ozone, allowing ground-based observations and communications. In contrast, regions like far-infrared and portions of ultraviolet are heavily absorbed, necessitating space-based detection and reinforcing separate regional identities.[36] These windows have shaped subdivisions, such as the radio band's breakdown into extremely low frequency (ELF, 3–30 Hz) to very high frequency (VHF, 30–300 MHz) for navigation and broadcasting.[37] Classifications have evolved from an initial focus on the optical regime (visible light) in the 18th century to a comprehensive spectrum by the early 20th century, driven by sequential discoveries. Prior to 1800, only visible light was recognized; William Herschel's 1800 detection of infrared and Johann Ritter's 1801 ultraviolet findings expanded the boundaries, followed by James Clerk Maxwell's 1867 theoretical prediction and Heinrich Hertz's 1887 experimental confirmation of radio waves. Wilhelm Röntgen's 1895 discovery of X-rays and Paul Villard's 1900 observation of gamma rays completed the full spectrum, shifting classifications from wavelength-centric optical divisions to energy-based, multi-region frameworks that accommodate technological advancements like radar and spectroscopy. Overlaps exist across boundaries—for example, the microwave-infrared transition around 1 mm or ultraviolet-visible at 400 nm—due to continuous wave properties, with sub-divisions (e.g., near-, mid-, and far-infrared based on thermal effects) providing finer granularity for specialized uses.[24][34]Radio Waves
Radio waves represent the longest-wavelength portion of the electromagnetic spectrum, characterized by wavelengths greater than 1 millimeter, corresponding to frequencies below 300 GHz. This range encompasses the practical radio frequency bands allocated for various uses, starting from very low frequency (VLF) at 3–30 kHz, low frequency (LF) at 30–300 kHz, medium frequency (MF) at 300 kHz–3 MHz, high frequency (HF) at 3–30 MHz, very high frequency (VHF) at 30–300 MHz, and ultra high frequency (UHF) at 300 MHz–3 GHz, with higher sub-bands extending up to 300 GHz.[38] These waves are generated by accelerating electric charges, typically through oscillating currents in antennas, where the frequency of oscillation determines the emitted wave's frequency.[39] Propagation of radio waves occurs primarily through ground waves, which follow the Earth's curvature for medium-range communication, and sky waves, which reflect off the ionosphere to enable long-distance transmission beyond the horizon. The ionosphere, a layer of ionized plasma in the upper atmosphere, refracts and reflects radio waves—particularly those in the HF band—due to free electrons interacting with the wave's electric field, allowing signals to bounce multiple times for global reach. Atmospheric effects, such as absorption in the lower ionosphere during daytime, influence propagation reliability, but nighttime conditions often enhance sky-wave efficiency for international broadcasting.[40][41].pdf) With photon energies on the order of 10^{-12} to 10^{-3} eV—far below the 13.6 eV threshold for ionizing atomic hydrogen—radio waves are non-ionizing and pose no risk of chemical bond disruption in biological tissues. Quantum effects are negligible at these low frequencies and energies, permitting a classical electromagnetic treatment using Maxwell's equations without invoking photon discreteness. Key applications include amplitude modulation (AM) radio broadcasting in the MF band for wide-area audio transmission and frequency modulation (FM) radio along with television signals in the VHF band, which offer higher fidelity due to reduced interference from atmospheric noise.[34][42][43][44]Microwaves
Microwaves constitute a segment of the electromagnetic spectrum characterized by wavelengths ranging from 1 millimeter to 1 meter, corresponding to frequencies between 300 GHz and 300 MHz.[45] This range positions microwaves as the higher-frequency extension of radio waves, enabling distinct propagation and interaction properties suitable for specialized technologies. Unlike longer radio waves, microwaves exhibit reduced diffraction and greater susceptibility to atmospheric attenuation, yet they maintain the ability to penetrate obscurants such as clouds, dust, and light precipitation.[45] Microwaves are generated primarily through vacuum tube devices like klystrons and magnetrons, which convert electrical energy into high-power electromagnetic oscillations. Klystrons operate via velocity modulation of an electron beam across multiple resonant cavities, achieving amplification with efficiencies up to 50% and gains exceeding 60 dB, making them ideal for high-power applications in accelerators and communications.[46] Magnetrons, in contrast, produce microwaves by inducing cycloidal electron paths in a crossed electric and magnetic field within resonant cavities, yielding compact, high-efficiency sources (around 80%) commonly used in radar and domestic appliances.[47] For propagation, microwaves rely on waveguides—hollow metallic structures that confine and guide waves with minimal loss—and parabolic dish antennas, which provide high directivity and beam focusing for point-to-point transmission over distances.[45] Microwaves interact strongly with water molecules due to excitation of their rotational modes, leading to significant absorption in both liquid and vapor forms. In the atmosphere, water vapor produces a prominent absorption line at approximately 22.235 GHz, while oxygen contributes lines around 60 GHz, creating regions of high attenuation.[48] These effects define "microwave windows"—frequency bands with relatively low absorption, such as 8–12 GHz and segments near 22 GHz, that facilitate transmission for remote sensing and communication by minimizing signal loss through the troposphere.[49] Key applications of microwaves leverage these properties for practical technologies. In microwave ovens, a magnetron generates 2.45 GHz waves that cause dielectric heating by inducing molecular rotation in water, fat, and sugars within food, achieving efficient volumetric cooking with penetration depths of several centimeters.[47] Radar systems utilize pulsed microwaves in bands like X-band (8–12 GHz) for high-resolution detection of objects, weather patterns, and terrain, benefiting from the waves' ability to penetrate atmospheric haze.[45] Satellite communications employ microwave frequencies in C-band (4–8 GHz) and Ku-band (12–18 GHz) for reliable transcontinental data relay, capitalizing on the microwave windows to ensure signal integrity.[45] Wireless networking standards like Wi-Fi operate in the 2.4 GHz and 5 GHz bands, enabling short-range, high-data-rate connectivity in unlicensed spectrum allocations.[50] At shorter wavelengths approaching 1 mm (millimeter waves), microwaves transition to quasi-optical behavior, where wave propagation mimics light rays more closely, allowing treatment via geometric optics approximations for beam focusing, reflection, and diffraction in free space. This shift enables advanced quasi-optical systems, such as Gaussian beam propagation in high-frequency devices, bridging microwave engineering with optical techniques.Infrared Radiation
Infrared radiation occupies the portion of the electromagnetic spectrum adjacent to visible light, with wavelengths ranging from approximately 700 nanometers to 1 millimeter, corresponding to frequencies between 430 terahertz and 300 gigahertz.[25] This band is subdivided into near-infrared (0.7–5 μm), mid-infrared (5–30 μm), and far-infrared (30 μm–1 mm), though exact boundaries vary slightly across conventions; these divisions reflect differences in interaction with matter and technological applications.[51] Unlike shorter wavelengths, infrared is invisible to the human eye and primarily manifests as thermal radiation from objects at everyday temperatures. Infrared emission is a key feature of blackbody radiation, where the peak wavelength of emission follows Wien's displacement law: λ_max T = 2898 μm·K, meaning warmer objects peak at shorter infrared wavelengths while cooler ones, like room-temperature surfaces, emit predominantly in the mid- to far-infrared.[52] Detection typically relies on thermal sensors such as bolometers, which measure temperature changes from absorbed radiation via resistance variations, or thermocouples, which generate voltage from heat-induced junctions; these methods enable sensitive imaging across the infrared range without requiring cryogenic cooling for many applications.[53] Infrared radiation interacts with matter by exciting molecular vibrations and rotations, leading to characteristic absorption bands—for instance, carbon dioxide absorbs strongly at 4.3 μm due to asymmetric stretching modes, as detailed in high-resolution spectroscopic studies identifying over 90% of lines in this region.[54] As non-ionizing radiation, infrared primarily causes heating by transferring energy to molecular bonds rather than ejecting electrons, making it suitable for non-destructive sensing.[51] Practical applications include thermal imaging for detecting heat signatures in firefighting or search-and-rescue operations, remote controls using near-infrared light-emitting diodes at around 940 nm to transmit signals, and astronomy via telescopes like NASA's Spitzer Space Telescope, which observed dust-enshrouded star-forming regions in the mid- to far-infrared from 2003 to 2020.[25] In Earth's energy balance, infrared plays a central role as the primary form of outgoing thermal radiation from the surface (peaking around 10 μm), with absorption by atmospheric gases like CO2 trapping about 5–6% of incoming solar energy and contributing to the natural greenhouse effect that raises global temperatures by roughly 30°C.[55]Visible Light
Visible light constitutes the portion of the electromagnetic spectrum that is detectable by the human eye, spanning wavelengths from approximately 400 to 700 nanometers (nm), corresponding to frequencies between about 430 and 750 terahertz (THz).[56] This narrow band is subdivided into colors based on wavelength, progressing from violet at the shorter end (around 400 nm) to red at the longer end (around 700 nm), with intermediate hues including blue, green, yellow, and orange.[6] The human visual system perceives this spectrum through three types of cone photoreceptor cells in the retina, each tuned to peak sensitivities at distinct wavelengths: short-wavelength cones (S-cones) at about 420 nm (blue-violet), medium-wavelength cones (M-cones) at around 530 nm (green), and long-wavelength cones (L-cones) at approximately 560 nm (yellow-red).[57] These cones enable trichromatic color vision, where perceived colors arise from the combined responses of the cones, following additive color mixing models such as RGB (red, green, blue) used in digital displays.[58] Key optical properties of visible light include refraction and dispersion, which occur when light passes through media like glass prisms, causing shorter wavelengths (violet) to bend more than longer ones (red) due to wavelength-dependent refractive indices.[59] This phenomenon, first systematically demonstrated by Isaac Newton in the late 17th century, separates white light into its spectral components, producing rainbows in natural settings like atmospheric water droplets or artificial prisms.[60] Additionally, interference effects arise in thin films, such as soap bubbles or oil slicks on water, where the superposition of reflected light waves from the film's top and bottom surfaces creates colorful patterns depending on the film's thickness relative to the wavelength.[61] Primary sources of visible light include the Sun, whose blackbody radiation at about 5800 K peaks in the green-yellow region around 500 nm, delivering roughly 40-50% of its energy in the visible range to Earth's surface.[62] Artificial sources encompass incandescent lamps, fluorescent tubes, light-emitting diodes (LEDs) that emit specific colors via semiconductor bandgaps, and lasers producing coherent monochromatic beams, such as helium-neon lasers at 632.8 nm (red).[63] Applications leverage these properties in photography for capturing spectral scenes, display technologies like LCD and OLED screens employing RGB backlights for color reproduction, and fiber optic communications transmitting visible wavelengths over short distances in multimode fibers.[56] Biologically, visible light plays a central role in human vision, enabling object recognition and environmental navigation, while specific wavelengths influence physiological processes such as circadian rhythms, where blue light (around 450-480 nm) suppresses melatonin production via intrinsically photosensitive retinal ganglion cells, helping synchronize sleep-wake cycles to day-night patterns.[64] In plants, visible light drives photosynthesis, with chlorophyll pigments absorbing primarily blue (430-450 nm) and red (640-680 nm) wavelengths to convert light energy into chemical bonds, sustaining global ecosystems.[6]Ultraviolet Radiation
Ultraviolet radiation occupies the portion of the electromagnetic spectrum with wavelengths ranging from 10 to 400 nanometers, corresponding to frequencies between 750 terahertz and 30 petahertz.[26][3] This band is subdivided into three regions based on wavelength: UVA from 315 to 400 nm, UVB from 280 to 315 nm, and UVC from 100 to 280 nm, with each exhibiting distinct interactions with matter due to increasing photon energy.[26] Photon energies in this range span ~3–124 eV. While UV is generally classified as non-ionizing radiation that primarily affects molecular bonds rather than ejecting inner-shell electrons, shorter wavelengths (below ~124 nm) can cause ionization.[34] The Earth's ozone layer plays a critical role in filtering UV radiation, absorbing nearly all UVC and most UVB while allowing UVA to pass through more readily.[65] This absorption protects surface life from the most energetic UV components, as UVC is completely blocked by ozone and atmospheric oxygen.[65] Artificial sources of UV radiation include electric arcs, which emit broad-spectrum UV through plasma excitation, and mercury vapor lamps, which produce discrete UV lines, particularly at 254 nm in low-pressure configurations for targeted applications.[66][67] UV radiation drives key photochemical reactions by providing photons energetic enough to break molecular bonds, such as in the formation of stratospheric ozone where ultraviolet light photodissociates oxygen molecules:\ce{O2 + h\nu -> O + O}
followed by recombination with another oxygen molecule to form ozone (\ce{O + O2 + M -> O3 + M}, where M is a third body).[68] This process exemplifies UV's role in initiating chain reactions that alter atmospheric composition without requiring ionization.[69] Practical applications leverage UV's reactivity for sterilization, where UVC disrupts microbial DNA to inactivate pathogens in air, water, and surfaces.[70] In fluorescence, UVA excites electrons in materials to produce visible emission, enabling uses in detection and imaging.[71] Tanning beds employ UVA to stimulate melanin production for cosmetic skin darkening.[72] In astronomy, space-based observatories like NASA's Galaxy Evolution Explorer (GALEX) have mapped UV emissions from young stars across the universe, revealing star formation histories inaccessible from ground-based telescopes due to atmospheric absorption.[73] Exposure to UV radiation can cause acute health effects, including sunburn from erythema induced by UVB and DNA damage such as pyrimidine dimer formation that impairs cellular replication.[74] While non-ionizing, prolonged UVA and UVB exposure contributes to cumulative skin damage, emphasizing the need for protective measures against overexposure.[74]