Infrared
Infrared radiation, often abbreviated as IR, is a form of electromagnetic radiation with wavelengths longer than those of visible red light but shorter than microwaves, typically spanning from approximately 700 nanometers (nm) to 1 millimeter (mm).[1] This places it immediately beyond the visible spectrum in the electromagnetic spectrum, where it is invisible to the human eye but commonly experienced as thermal energy or heat.[2] Discovered in 1800 by astronomer William Herschel through experiments dispersing sunlight with a prism and measuring temperature variations, infrared radiation is emitted by all objects with temperatures above absolute zero, following principles of blackbody radiation where emission intensity increases with temperature.[1] Infrared is subdivided into regions such as near-infrared (0.7–1.4 μm), short-wavelength infrared (1.4–3 μm), mid-infrared (3–8 μm), long-wavelength infrared (8–15 μm), and far-infrared (15–1,000 μm); detailed schemes are discussed in the Spectral Regions section. These regions have distinct properties and applications due to varying photon energies from about 1.2 meV to 1.7 eV. Its longer wavelengths allow it to penetrate dust and dense gases with less scattering than visible light, making it valuable for observing obscured cosmic phenomena, while its thermal characteristics enable detection of heat signatures from warm bodies. Natural sources include the Sun's emission peaking in the near-infrared, Earth's atmospheric and surface heat in the thermal infrared, and cool celestial objects like planets and star-forming regions.[2] Artificial sources encompass heat lamps emitting broadly from 500 nm to 3,000 nm, as well as electronic devices like remote controls operating in the near-infrared around 940 nm.[2] Key applications of infrared radiation span multiple fields, including thermal imaging for night vision and firefighting via sensors detecting 8–15 μm wavelengths, astronomy through telescopes like the James Webb Space Telescope that reveal planet-forming disks hidden by dust, and Earth observation satellites such as MODIS for monitoring wildfires and vegetation health. In medicine, low-level infrared therapy utilizes near-infrared wavelengths (760–1,000 nm) for photobiomodulation to promote tissue repair and reduce inflammation, while far-infrared (>15 μm) is explored for its potential biological effects like improved circulation.[4] Additionally, infrared spectroscopy identifies molecular structures by measuring absorption in the mid- to far-infrared, and it powers consumer technologies from motion sensors to fiber-optic communications in the near-infrared band.[1] Despite its benefits, prolonged exposure to intense infrared can cause thermal burns, highlighting the need for safety protocols in industrial and medical uses.[5]Fundamentals
Definition and Electromagnetic Spectrum
Infrared radiation, often abbreviated as IR, is a form of electromagnetic radiation characterized by wavelengths ranging from approximately 700 nanometers (nm) to 1 millimeter (mm). This corresponds to a frequency range of about 430 terahertz (THz) to 300 gigahertz (GHz).[6] These parameters place infrared just beyond the red end of the visible spectrum, where human eyes perceive light up to around 700 nm.[7] Within the broader electromagnetic spectrum, infrared occupies the band between visible light—with its shorter wavelengths from about 400 nm to 700 nm—and microwaves, which extend from 1 mm to 1 meter in wavelength. This positioning means infrared has longer wavelengths and lower frequencies than visible light but higher frequencies than microwaves, influencing its interactions with matter, such as penetration and absorption properties. Infrared radiation is emitted by all objects above absolute zero due to thermal motion, though this thermal aspect is explored further elsewhere.[6] As with all electromagnetic radiation, infrared demonstrates wave-particle duality, manifesting as propagating waves or discrete packets of energy known as photons. The energy E of an individual infrared photon is determined by the formulaE = h \nu,
where h is Planck's constant ($6.626 \times 10^{-34} J·s) and \nu is the photon's frequency. This relationship underscores the quantized nature of infrared energy, with photon energies typically ranging from about 1.24 millielectronvolts (meV) to 1.7 electronvolts (eV) across the infrared band.[9] Infrared wavelengths are typically measured in nanometers (nm) for near-infrared regions or micrometers (μm) for longer wavelengths, with 1 μm equaling 1000 nm. In infrared spectroscopy, the reciprocal of wavelength, known as wavenumber and expressed in inverse centimeters (cm⁻¹), is commonly used; for example, the infrared range spans roughly 14,000 cm⁻¹ to 10 cm⁻¹. This unit facilitates analysis of molecular vibrations, as higher wavenumbers correspond to higher energies.
Physical Properties
Infrared radiation, as a form of electromagnetic waves, propagates through vacuum at the speed of light, c = 3 \times 10^8 m/s.[10] In various media, its speed is reduced according to v = c / n, where n is the refractive index of the medium, which for infrared wavelengths typically ranges from about 1.3 to 4 depending on the material, such as glasses or semiconductors used in optics.[11] This velocity reduction leads to phenomena like refraction when infrared waves pass through interfaces between media, enabling applications in lenses and waveguides.[12] Infrared radiation interacts with matter primarily through absorption, reflection, transmission, and scattering processes.[13] Absorption occurs when infrared photons excite molecular vibrations, converting energy into heat; reflection and transmission depend on the material's surface properties and refractive index contrast; while scattering redirects waves without significant energy loss, often via Rayleigh or Mie mechanisms for small particles.[14] In the Earth's atmosphere, these interactions are pronounced, with strong absorption bands due to water vapor (particularly in the 2.7 μm and 6.3 μm regions) and carbon dioxide (around 4.3 μm and 15 μm), which attenuate infrared transmission over long paths and contribute to the greenhouse effect.[15] Atmospheric aerosols further enhance scattering and absorption, reducing visibility in the infrared spectrum.[14] A key aspect of infrared-matter interaction is penetration depth, which varies by wavelength and material. Near-infrared (0.7–1.4 μm) penetrates biological tissues deeper, up to 10–15 mm in models like bovine tissue, due to lower absorption by water and hemoglobin compared to longer wavelengths.[16] In contrast, far-infrared (above 3 μm) is absorbed more superficially, up to several centimeters (e.g., ~4 cm) in skin, as it strongly excites water molecule vibrations, enabling heating effects in deeper tissues while primarily interacting with surface layers.[17] Infrared radiation exhibits polarization and coherence properties inherent to electromagnetic waves. Polarization describes the orientation of the electric field vector, which can be linear, circular, or elliptical; natural sources like thermal emitters produce unpolarized infrared, but it can be polarized through reflection, scattering, or devices like wire-grid polarizers.[18] Coherence refers to the phase correlation of waves: thermal infrared sources, such as blackbodies, are typically incoherent with short coherence lengths (on the order of micrometers), leading to diffuse emission, whereas infrared lasers generate highly coherent radiation with long coherence lengths (centimeters to meters), enabling interferometry and precise imaging.[19][20]Spectral Regions
Common Subdivision Schemes
The infrared spectrum is commonly subdivided into three primary regions—near-infrared (NIR), mid-infrared (MIR), and far-infrared (FIR)—based on wavelength ranges that align with distinct physical interactions and technological capabilities. The near-infrared region spans approximately 0.7 to 1.4 μm, overlapping with the visible spectrum and enabling applications like fiber optics and night vision due to its proximity to visible light.[21] The mid-infrared region extends from 1.4 to 15 μm, often further divided into short-wavelength infrared (1.4–3 μm) and longer mid-wave (3–15 μm) sub-bands to reflect varying absorption characteristics.[22] The far-infrared region covers 15 to 1000 μm, where thermal emission dominates and wavelengths approach millimeter scales.[21] These divisions arise from key physical and practical considerations, including the excitation of molecular vibrations, which are most prominently observed in the mid-infrared where fundamental stretching and bending modes occur, providing a basis for spectroscopic identification of chemical bonds.[23] Atmospheric transmission windows also influence the boundaries, with relatively clear paths in the near-infrared (around 0.7–1.1 μm) and mid-infrared (3–5 μm and 8–14 μm) allowing for effective ground-based observations and remote sensing, while stronger absorption by water vapor and CO₂ limits transmission elsewhere.[24] Detector sensitivities further justify the scheme, as silicon-based photodiodes excel in the near-infrared, mercury cadmium telluride (MCT) detectors perform optimally in the mid-infrared, and microbolometers or cryogenic sensors are required for the far-infrared due to lower photon energies.[21] In spectroscopic contexts, these wavelength ranges correspond to specific wavenumber intervals, with the mid-infrared commonly expressed as 4000–400 cm⁻¹ to encompass the functional group and fingerprint regions for molecular analysis.[23] Boundaries exhibit overlaps and variations across disciplines; for instance, optics often uses 0.78–3 μm for near- to mid-infrared transitions to align with laser sources, while spectroscopy may extend near-infrared to 2.5 μm to include overtone vibrations.[21] Such inconsistencies stem from historical conventions and application-specific needs, like eye safety standards (e.g., IR-A for near-infrared) versus thermal imaging bands.[22]Specialized Division Schemes
The International Commission on Illumination (CIE) defines infrared radiation as optical radiation with wavelengths longer than the visible spectrum, starting at 780 nm where human visual response diminishes, though sensation may extend slightly beyond for intense sources; it further subdivides IR into IR-A (780 nm to 1.4 μm), IR-B (1.4 to 3 μm), and IR-C (3 μm to 1 mm) to align with physiological effects on the eye.[25] This scheme prioritizes boundaries tied to human perception and biological interaction rather than uniform wavelength intervals. The ISO 20473:2007 standard, developed for optics and photonics applications excluding lighting and telecommunications, divides infrared into short-wave (near-infrared, 0.78–3 μm), mid-wave (3–50 μm), and long-wave (50–1000 μm) regions to standardize spectral band descriptions across technical fields.[17] Unlike broader schemes, this emphasizes practical delimitation for instrument calibration and material testing, with mid-wave encompassing vibrational spectroscopy bands and long-wave covering thermal emissions. In astronomy, infrared is segmented into near-IR (0.8–5 μm), mid-IR (5–30 μm), and far-IR (30–1000 μm) to match observational challenges like atmospheric absorption and telescope capabilities; the far-IR regime is particularly vital for detecting thermal emissions from cosmic dust grains, which absorb ultraviolet-visible starlight and re-emit at these longer wavelengths, revealing interstellar structures such as molecular clouds and protostellar envelopes.[26] Telecommunications employs fine-grained near-IR bands optimized for low-loss propagation in silica fiber optics, including the O-band (1260–1360 nm) for short-haul metropolitan networks and the C-band (1530–1565 nm) for long-haul amplification via erbium-doped fibers, enabling high-capacity data transmission with minimal signal attenuation.[27] Infrared divisions based on sensor response align with detector material sensitivities, such as indium antimonide (InSb) photodiodes effective in the mid-IR (1–5.5 μm) for high-speed, cooled applications like thermal imaging, while mercury cadmium telluride (HgCdTe) extends to long-wave IR (up to 15 μm) and bolometers handle far-IR beyond 20 μm via thermal absorption.[28] These material-driven boundaries facilitate selection for specific detection tasks, prioritizing quantum efficiency over arbitrary wavelength cuts.Thermal Radiation
Infrared as Heat Radiation
Infrared radiation serves as the primary mechanism for thermal emission from objects at room temperature, where the peak wavelength of emission falls within the long-wavelength infrared range around 10 μm, as determined by Wien's displacement law. This law states that the wavelength at which a blackbody emits the maximum intensity of radiation, λ_max, is inversely proportional to its absolute temperature T, given by the formula λ_max T = 2.897 × 10^{-3} m·K. For typical room temperatures of approximately 300 K, this yields λ_max ≈ 9.7 μm, placing the strongest emission in the infrared spectrum and explaining why everyday objects like walls or furniture radiate heat primarily as infrared waves.[29] Kirchhoff's law of thermal radiation further elucidates this process by establishing that, at thermal equilibrium and for a given wavelength, the emissivity ε of a surface—which measures its efficiency in emitting radiation—equals its absorptivity α, the fraction of incident radiation it absorbs. Thus, surfaces that are good absorbers of infrared radiation, such as dark or matte materials, are also efficient emitters of thermal infrared at the same wavelengths, ensuring a balance in radiative heat exchange. This principle underlies why blackened surfaces, which absorb nearly all incident radiation (α ≈ 1), emit strongly when heated, while reflective surfaces like polished metals emit poorly.[30] A practical example is the human body, which at a normal temperature of about 310 K emits peak infrared radiation in the long-wavelength infrared band at approximately 9.3 μm, contributing to sensible heat loss in environments. In daily life, infrared facilitates heat transfer through radiation, such as the warmth felt from sunlight on the skin—where solar infrared accounts for much of the heating effect—or the radiative cooling of a hot cup of coffee to the surrounding air, independent of conduction or convection.[31][32] It is important to distinguish thermal infrared sources, which arise from the random thermal motion of particles in matter and follow blackbody-like spectra dependent on temperature, from non-thermal sources that produce infrared without significant heat generation. Non-thermal examples include infrared lasers or light-emitting diodes (LEDs) used in remote controls, which emit coherent or directed beams at specific wavelengths like 940 nm through electronic excitation rather than thermal processes.[7]Blackbody Emission
A blackbody is defined as an idealized physical body that absorbs all incident electromagnetic radiation, regardless of frequency or angle of incidence, and re-emits energy solely based on its temperature, with no reflection or transmission.[33] This perfect absorber also serves as a perfect emitter, providing a theoretical benchmark for thermal radiation spectra.[34] The spectral radiance of blackbody radiation is described by Planck's law, which quantifies the power emitted per unit area per unit solid angle per unit frequency as a function of wavelength \lambda and temperature T: B(\lambda, T) = \frac{2hc^2}{\lambda^5} \frac{1}{e^{hc / \lambda k T} - 1}, where h is Planck's constant, c is the speed of light, and k is Boltzmann's constant.[34] This formula reveals that the emission spectrum peaks at a wavelength that shifts with temperature, with significant infrared contributions dominating at lower temperatures due to the exponential term in the denominator.[35] Wien's displacement law further characterizes this peak shift, stating that the wavelength \lambda_{\max} at which the spectral radiance is maximum satisfies \lambda_{\max} T = b, where b \approx 2898 \, \mu\mathrm{m \cdot K} is Wien's displacement constant.[35] For terrestrial temperatures around 300 K, \lambda_{\max} falls in the long-wavelength infrared range near 10 \mum, explaining why blackbody emission from everyday objects is predominantly infrared.[36] The total radiant power emitted by a blackbody is governed by the Stefan-Boltzmann law, P = \sigma A T^4, where \sigma = 5.670 \times 10^{-8} \, \mathrm{W \cdot m^{-2} \cdot K^{-4}} is the Stefan-Boltzmann constant and A is the surface area.[37] At temperatures below 1000 K, the majority of this power is emitted in the infrared spectrum, as shorter wavelengths become negligible.[36]Detection and Measurement
Infrared Detectors
Infrared detectors convert incident infrared radiation into measurable electrical signals, broadly categorized into thermal detectors, which respond to the heat generated by absorbed photons, and quantum detectors, which rely on photon-induced electronic transitions. Thermal detectors operate by sensing temperature changes caused by radiation absorption and can function across all infrared spectral bands without wavelength selectivity, making them suitable for broadband detection. Quantum detectors, in contrast, exhibit wavelength-specific responses based on material bandgaps and typically offer higher sensitivity but often require cooling to minimize thermal noise.[38][39] Thermal detectors include bolometers, which measure radiation through changes in electrical resistance due to heating of a sensitive element, such as a thin film of vanadium oxide or amorphous silicon. The resistance variation is proportional to the temperature rise from absorbed infrared power, enabling room-temperature operation with responsivities on the order of 10^4 to 10^5 V/W in microbolometer arrays. Pyroelectric detectors, another type of thermal device, generate a voltage in response to rapid temperature changes via the pyroelectric effect in materials like lithium tantalate or PZT ceramics, producing charge proportional to the rate of incident radiation flux. These detectors are inherently AC-coupled, requiring modulation of the input signal, and achieve noise equivalent powers (NEPs) as low as 10^{-9} W/√Hz at room temperature. Both bolometer and pyroelectric types are uncooled and versatile for near- to far-infrared detection, though they exhibit slower response times (milliseconds) compared to quantum alternatives.[40][41][42] Quantum detectors, such as photodiodes, operate by absorbing photons to excite charge carriers across a bandgap, with materials like mercury cadmium telluride (HgCdTe) commonly used for mid-infrared detection due to its tunable bandgap (0.1–1.5 eV) via composition control. In photovoltaic mode, the photodiode generates a voltage without external bias, relying on the built-in electric field to separate electron-hole pairs, which suits low-noise, power-limited applications. Photoconductive mode applies reverse bias to increase carrier collection speed and gain, though it introduces higher dark current; quantum efficiency η, defined as \eta = \frac{\text{number of charge carriers generated}}{\text{number of incident photons}}, quantifies the conversion efficiency, often exceeding 70% for optimized HgCdTe devices in the 3–5 μm band. These detectors provide high speed (nanoseconds) and spectral selectivity but suffer from thermal generation noise at longer wavelengths.[43][44][45] For mid- and far-infrared quantum detection, cryogenic cooling is essential to reduce thermal excitation of carriers, which otherwise overwhelms the signal; typical operating temperatures are 77 K using liquid nitrogen or below 50 K with mechanical coolers for HgCdTe or InSb devices, achieving background-limited performance with dark currents reduced by orders of magnitude. Uncooled operation is feasible for short-wave infrared but degrades detectivity in longer bands due to increased noise. A key figure of merit for comparing detector performance is the specific detectivity D^*, given by D^* = \frac{\sqrt{A \Delta f}}{\text{NEP}}, where A is the detector area, Δf is the electrical bandwidth, and NEP is the noise equivalent power (the incident power yielding a signal-to-noise ratio of 1 in 1 Hz bandwidth, in W/√Hz); D^* normalizes for size and speed, with values around 10^{10} cm √Hz/W for room-temperature thermal detectors and up to 10^{12} cm √Hz/W for cooled quantum types.[46][47][48]Spectroscopy Techniques
Infrared spectroscopy techniques enable the analysis of materials by measuring their interaction with infrared radiation, primarily through absorption, emission, or reflection of light in the mid-infrared region (approximately 4000–400 cm⁻¹). These methods exploit the fact that molecules absorb IR radiation at frequencies corresponding to their vibrational transitions, producing spectra that serve as unique signatures for chemical identification and quantification. Instrumentation typically involves a broadband IR source, a sample interface, and detectors to record the modulated signal, with data processed to yield absorbance or transmittance spectra.[49] Fourier-transform infrared (FTIR) spectroscopy is the predominant technique, utilizing a Michelson interferometer to achieve high-resolution spectra across a wide wavelength range simultaneously. In the Michelson setup, incoming IR radiation is split by a beam splitter into two paths: one reflects off a fixed mirror, while the other travels to a movable mirror that scans along the optical axis, creating a variable path difference. The recombined beams interfere, producing an interferogram—a time-domain signal that encodes the spectral information—which is then converted to a frequency-domain spectrum via Fourier transform. This approach offers advantages over dispersive methods, including higher signal-to-noise ratios and faster acquisition times. The spectral resolution in FTIR, denoted as Δν (in cm⁻¹), is determined by the maximum mirror displacement L (in cm), following the relation Δν = 1/(2L), where 2L represents the maximum optical path difference. For example, a maximum path difference of 2 cm yields a resolution of 0.5 cm⁻¹, allowing distinction of closely spaced absorption bands.[49][49] Absorption spectroscopy, often performed in transmission mode, quantifies how much IR light is absorbed by a sample as it passes through, providing direct insight into molecular concentrations. The fundamental principle is the Beer-Lambert law, which states that absorbance A is linearly proportional to the analyte concentration c (in mol/L), the path length l (in cm), and the molar absorptivity ε (in L mol⁻¹ cm⁻¹) at a specific wavelength:A = \varepsilon c l
This law holds for dilute solutions or gases where interactions are minimal, enabling quantitative analysis such as determining protein concentrations in biochemical samples from characteristic amide band intensities. In practice, samples are prepared as thin films, pellets (e.g., KBr matrices for solids), or gas cells to ensure sufficient transmission without saturation. Deviations occur at high concentrations due to non-ideal behaviors, but the law remains foundational for calibration curves in analytical applications.[50][50] Emission and reflection modes extend IR spectroscopy to samples unsuitable for transmission, such as opaque solids or viscous liquids, by measuring emitted or reflected radiation instead of transmitted light. In emission spectroscopy, thermally excited samples (e.g., at elevated temperatures) radiate IR light characteristic of their vibrational modes, useful for studying hot gases or surfaces. Reflection techniques, particularly attenuated total reflectance (ATR), facilitate analysis without extensive preparation by exploiting total internal reflection at a high-refractive-index crystal (e.g., diamond or ZnSe) in contact with the sample. The evanescent wave penetrates ~1–5 μm into the sample, interacting with it and attenuating based on absorption; the reflected beam is then analyzed via FTIR. ATR is ideal for solids and liquids, requiring only a small amount of material pressed against the crystal, and spectra are corrected for the wavelength-dependent penetration depth to mimic transmission data. This mode minimizes artifacts from sample thickness and is widely used for quality control in polymers and pharmaceuticals. A key aspect of spectral interpretation in mid-IR spectroscopy is the functional group region (4000–1500 cm⁻¹), featuring characteristic absorptions from functional groups, such as O-H stretches near 3500 cm⁻¹ or C=O stretches around 1700 cm⁻¹. The fingerprint region (1500–400 cm⁻¹), where complex combinations of stretching and bending vibrations produce unique patterns for molecular identification, allows differentiation of isomers or contaminants by matching against reference libraries. Unlike the higher-wavenumber region, the fingerprint area is more skeletal-specific and provides a holistic "signature" for compound verification, often confirmed by overlaying experimental spectra with standards in database searches.[51]