Fact-checked by Grok 2 weeks ago

Spectral band

A spectral band is a well-defined, continuous or range of or frequencies within the , typically used to isolate and measure specific portions of for analysis in scientific and technical applications. These bands are characterized by a central and a , enabling precise segmentation of the spectrum for various purposes. In and , spectral bands form the basis of multispectral and systems, where sensors capture reflected or emitted energy from Earth's surface in discrete intervals to identify materials, health, and environmental changes based on unique spectral signatures. For instance, the satellite employs 11 spectral bands across the visible, near-infrared, shortwave infrared, and thermal infrared regions, including Band 1 (coastal : 0.43–0.45 µm at 30 m resolution) for atmospheric correction and Band 5 (near-infrared: 0.85–0.88 µm at 30 m resolution) for monitoring. Band 4 (red: 0.64–0.67 µm) and Band 3 (green: 0.53–0.59 µm) contribute to natural color composites, while thermal bands like Band 10 (10.60–11.19 µm at 100 m resolution) assess surface temperature. These bands allow for applications such as classification, crop monitoring, and disaster assessment by exploiting differences in how objects reflect or absorb at specific wavelengths. In , spectral bands refer to regions where , , or of occurs, revealing molecular and structures through characteristic patterns like bands in spectra. Instruments in this field use narrow spectral bands to resolve fine details, such as vibrational transitions in molecules, aiding in chemical and material . In astronomy and radio communications, spectral bands denote designated frequency ranges, such as the S-band (2–4 GHz) used for satellite telemetry or X-band (8–12 GHz) for observations, facilitating interference-free signal transmission and celestial object studies. Overall, the design and selection of spectral bands are critical for optimizing , , and data utility across these domains, with ongoing advancements in sensor technology enabling finer bandwidths for more detailed .

Fundamentals

Definition

A spectral band is a contiguous range of wavelengths, frequencies, or wavenumbers within an where particular physical processes, such as absorption or emission of radiation, take place or are observed. These bands represent regions of enhanced or diminished intensity due to interactions between matter and , commonly observed in spectroscopic studies of atoms, molecules, and materials. Spectral bands are typically denoted using units of wavelength (e.g., nanometers [nm] or micrometers [μm]), frequency (e.g., hertz [Hz] or terahertz [THz]), or wavenumber (e.g., inverse centimeters [cm⁻¹]). For instance, the visible light band spans approximately 400–700 nm, corresponding to frequencies around 430–750 THz and wavenumbers of about 14,000–25,000 cm⁻¹. Unlike discrete spectral lines, which appear as narrow peaks from individual quantum transitions in atomic spectra, spectral bands are broader features arising from the overlap of numerous closely spaced lines, often due to rotational and vibrational energy levels in molecular systems. This broadening distinguishes bands as extended regions rather than isolated features, reflecting the complexity of multi-level quantum interactions. The concept of spectral bands originated in early 20th-century , as researchers applied emerging to interpret molecular emission and absorption patterns. Building on Niels Bohr's 1913 atomic model, which explained discrete line spectra in , scientists extended these ideas to molecular systems, developing the quantum theory of band spectra by the 1920s. This foundational work marked a shift toward understanding continuous spectral regions in polyatomic structures.

Characteristics

Spectral bands exhibit several fundamental physical properties that enable their quantification and in spectroscopic measurements. The of a band, which reflects the strength of the underlying transition, can be assessed either as the peak height—the maximum or deviation—or as the integrated area under the curve, providing a measure of the total transition probability. The central position of the band is defined by its or , corresponding to the energy difference between quantum states involved in the transition. Band width is typically characterized by the (FWHM), representing the range over which the drops to half its peak value, often on the order of picometers to nanometers depending on the system. describes deviations from a symmetric profile, arising from factors like overlapping transitions or instrumental effects, and is quantified by ratios such as the parameter or tail-to-peak ratios. The observed characteristics of spectral bands are significantly influenced by the resolution of the measuring instrument. Instrumental resolution determines the ability to distinguish fine spectral features, with poorer resolution leading to apparent broadening and merging of bands that would otherwise appear distinct. The resolving power R is formally defined as R = \frac{\lambda}{\Delta \lambda}, where \lambda is the central and \Delta \lambda is the minimum resolvable separation, typically achieving values from hundreds to tens of thousands in modern spectrometers. The distribution and intensity of spectral bands are governed by quantum mechanical principles, particularly the transition probabilities between energy levels. In atomic and molecular systems, the oscillator strength f, a dimensionless quantity ranging from near zero for forbidden transitions to around 1 for fully allowed ones, directly scales with the band's integrated intensity and encapsulates the likelihood of or . Environmental conditions play a crucial role in modifying band properties through broadening mechanisms. induces from the thermal motion of particles, resulting in a Gaussian profile with FWHM given by \Delta \nu_D = \frac{\nu_0}{c} \sqrt{\frac{8 k T \ln 2}{m}}, where \nu_0 is the central frequency, c is the , k is Boltzmann's constant, T is , and m is the of the emitting species; this effect increases with higher temperatures and lighter particles. Pressure contributes to collisional or Lorentzian broadening via interactions that shorten excited-state lifetimes, while the surrounding medium can introduce additional shifts or widths through variations or . These influences are particularly relevant in contexts like and , where band profiles inform about molecular environments.

In Spectroscopy and Optics

Absorption and Emission Bands

Absorption bands occur in regions of the where molecules or atoms absorb photons, leading to excitation of electrons from ground to higher energy states or to vibrational modes within molecules. This is characteristic of specific substances and provides insight into their electronic and vibrational structures. For instance, in ultraviolet-visible (UV-Vis) , organic dyes like exhibit distinct absorption bands; shows strong absorptions at approximately 430 (blue) and 662 (red), corresponding to π → π* transitions in its porphyrin ring. Emission bands arise when excited species relax to lower energy states, releasing photons in processes such as (rapid emission from singlet states) or (delayed emission from triplet states). These bands typically appear at longer wavelengths than absorption bands due to energy loss through vibrational relaxation. A key feature is the , defined as the difference between the absorption and emission maxima, which for fluorescent dyes like fluorescein is around 20-30 nm and arises from reorganization of the solvent shell around the excited molecule. Quantitative analysis of absorption bands relies on the Beer-Lambert law, which states that the absorbance A of a solution is directly proportional to the concentration c of the absorbing species, the molar absorptivity \epsilon (a constant specific to the substance and wavelength), and the path length l through the sample: A = \epsilon c l This law enables determination of unknown concentrations in , such as measuring dye concentrations in solutions, provided the solution is dilute to avoid deviations from linearity. In () spectroscopy, absorption bands often involve coupled rotational-vibrational transitions, forming complex band envelopes. For diatomic or linear polyatomic molecules, these spectra display P, Q, and R branches: the P branch corresponds to transitions where the rotational quantum number J decreases (\Delta J = -1), appearing at lower wavenumbers; the Q branch (\Delta J = 0) is sometimes present in perpendicular transitions of polyatomics, forming a central peak; and the R branch (\Delta J = +1) appears at higher wavenumbers. These branches envelope the pure vibrational frequency, with spacing influenced by the rotational constant B, as seen in the HCl fundamental vibration around 2886 cm⁻¹.

Band Shapes and Widths

Spectral bands in spectroscopy often exhibit characteristic shapes determined by underlying physical processes, with the most common profiles being Gaussian, Lorentzian, and Voigt. The Gaussian profile arises primarily from Doppler broadening due to the thermal motion of emitting or absorbing particles, following a Maxwell-Boltzmann velocity distribution. Its intensity as a function of frequency detuning \Delta \nu = \nu - \nu_0 is given by I(\Delta \nu) = I_0 \exp\left(-\frac{(\Delta \nu)^2}{2\sigma^2}\right), where \sigma is the standard deviation, and the full width at half maximum (FWHM) is \Delta \nu = \frac{2\nu_0}{c} \sqrt{2 \ln 2 \frac{kT}{m}}, with \nu_0 the central frequency, c the speed of light, k Boltzmann's constant, T temperature, and m the particle mass. In contrast, the Lorentzian profile results from natural or lifetime broadening, where the finite lifetime \tau of excited states leads to an uncertainty in energy levels. The profile is I(\Delta \nu) = I_0 \frac{(\Delta \nu_{1/2}/2)^2}{(\Delta \nu)^2 + (\Delta \nu_{1/2}/2)^2}, with FWHM \Delta \nu = \frac{1}{2\pi \tau}, derived from the Fourier transform of an exponentially decaying wavefunction. Pressure or collisional broadening also produces Lorentzian shapes by shortening the effective lifetime through intermolecular interactions. The represents the of Gaussian and components, accounting for combined Doppler and lifetime/pressure effects, and is prevalent in real spectra under atmospheric or conditions. It lacks a simple analytical form but can be approximated or computed numerically, with the overall FWHM depending on the relative contributions of the two profiles. Broadening mechanisms fundamentally shape these profiles. Natural broadening stems from the , \Delta E \Delta t \geq \hbar/2, where the energy uncertainty \Delta E = h \Delta \nu arises from the finite excited-state lifetime \Delta t = \tau. Pressure broadening occurs via collisions that interrupt wavefunction , scaling with gas density and leading to Lorentzian wings. Broadening is classified as homogeneous if all molecules experience the same shift and width (e.g., natural or collisional), resulting in uniform Lorentzian profiles, or inhomogeneous if subpopulations differ (e.g., Doppler due to variations), yielding Gaussian envelopes. To analyze overlapping bands, techniques separate individual profiles from composite . Least-squares fitting minimizes the difference between observed data and a model sum of profiles (e.g., Voigt functions), iteratively adjusting parameters like central , width, and amplitude while accounting for . This method is widely used in high-resolution to resolve subtle features. In () , vibrational band envelopes exemplify these concepts, arising from rovibrational transitions where rotational sublevels broaden the vibrational into a P-, Q-, and R-branch . For diatomic molecules like HCl, the envelope approximates a Gaussian or Voigt shape at due to dominant Doppler and effects, with the central Q-branch modulated by the rotational .

In Nuclear Physics

Energy Level Groupings

In nuclear physics, spectral bands manifest as clusters of closely spaced energy levels in atomic nuclei, where states with similar quantum configurations have excitation energies differing by keV to MeV scales. These groupings typically arise from shell model orbitals filled in specific ways, leading to collective behaviors in deformed nuclei, such as rotational sequences where angular momentum projections align to form coherent structures. Such bands form through the interplay of single-particle motions and collective excitations, often in nuclei distant from closed shells. In even-even nuclei, the ground-state rotational band exemplifies this, comprising levels with even spin values (I = 0^+, 2^+, 4^+, ...) built on a deformed intrinsic state. The energies follow the rigid rotor approximation, given by E_I = \frac{\hbar^2}{2 \mathcal{I}} I(I+1), where \mathcal{I} is the nuclear moment of inertia, reflecting the deformed charge distribution; for instance, in ^{154}Sm, the 2^+ state lies at 82 keV, with subsequent levels scaling accordingly to reveal the band's characteristic I(I+1) progression. This arises from nucleons in high-j orbitals contributing to deformation, as predicted by the unified model combining shell structure with collective rotation. These bands are observed primarily via gamma-ray spectroscopy, where de-excitation cascades produce sequences of stretched E2 transitions mapping the level spacings. In thermal neutron capture reactions like (n,\gamma), compound nuclei decay through such cascades, yielding spectra that delineate band structures; early measurements at facilities like the Oak Ridge High Flux Isotope Reactor (HFIR) resolved these in rare-earth isotopes, confirming rotational patterns up to I ≈ 10\hbar. The identification of these energy level groupings traces to the mid-1950s, evolving from the independent shell model proposals by and J. Hans D. Jensen in 1949, which explained and single-particle levels, to the collective rotational framework advanced by , Ben R. Mottelson, and Leo James Rainwater. Their unified model, integrating deformation with shell effects, was experimentally validated through quadrupole moment and transition rate studies, earning the 1975 .

Quantum Number Correlations

In deformed nuclei, rotational spectral bands are characterized by specific quantum numbers that correlate the angular momentum states within the band. The total spin quantum number I describes the magnitude of the nuclear , while the \pi (+ or -) indicates the behavior of the wave function under spatial inversion. The projection quantum number K, which represents the component of along the intrinsic symmetry axis of the deformed nucleus, serves as a primary label for band classification and distinguishes bands with the same I but different configurations. These s arise from the collective rotational model, where the nucleus behaves as a rigid or irrotational body, and their correlations ensure that band members share the same K^\pi at the bandhead, with I increasing in steps of 1 or 2. Bands are classified according to K values, reflecting different intrinsic deformations or vibrational excitations. The ground-state band typically has K = 0^+, corresponding to the lowest-energy configuration with . Excited K = 0^+ bands, known as bands, arise from oscillations along the symmetry axis, while gamma bands feature K = 2^+ and stem from transverse vibrations that break the . These classifications highlight how K correlates with the band's intrinsic structure, with \pi often conserved within the band due to the even nature of modes. Transitions within and between bands obey selection rules derived from the of the electromagnetic interaction. For intraband electric (E2) transitions, which dominate the decay patterns, the change in is \Delta I = 0, \pm 1, \pm 2 (excluding $0 \to 0), with no change, but consecutive intraband steps are stretched \Delta I = 2. Interband E2 transitions, such as those linking gamma to bands, favor \Delta I = 0, \pm 1 due to the \Delta K = 2 requirement. The rotational character is further confirmed by the \mathcal{I}, which signatures the band through the formula E(I) = \frac{\hbar^2}{2 \mathcal{I}} \left[ I(I+1) - K^2 \right] + E_0, where E_0 is the bandhead energy; variations in \mathcal{I} reflect the nuclear deformation and distinguish rotational from vibrational sequences. Isotopic variations in these correlations arise from changes in neutron and proton numbers, which alter the nuclear deformation and thus shift band positions and \mathcal{I} values. In the rare-earth region, for instance, the dysprosium isotopes near A = 160 exhibit systematic evolution: in ^{160}\mathrm{Dy} (with 94 neutrons and 66 protons), the ground and gamma bands display pronounced rotational features with \mathcal{I} increasing toward midshell due to enhanced quadrupole deformation from added neutrons filling deformed orbitals. Neighboring isotopes like ^{158}\mathrm{Dy} and ^{162}\mathrm{Dy} show contracted band spacings and altered K assignments as the neutron excess modifies the potential energy surface. Experimental identification of these bands and their assignments relies on reactions that selectively populate specific states. excitation, using heavy-ion beams at sub-barrier energies, excites low-spin members of rotational bands via absorption, allowing measurement of B(\mathrm{E2}) values to confirm K^\pi and \mathcal{I}. reactions, such as one-nucleon pickup or stripping with light ions, populate higher-spin states and enable angular distribution analysis to determine I and through Doppler-shift attenuation. These methods have been pivotal in mapping band structures in deformed nuclei, linking observed spectra to theoretical predictions without significant fragmentation.

In Remote Sensing and Imaging

Multispectral Applications

in involves capturing in a limited number of discrete, predefined bands, typically ranging from 3 to 15 narrow wavelength intervals, to enable the discrimination of materials based on their unique or emission signatures across these bands. This approach contrasts with panchromatic imaging by providing enhanced , allowing for the separation of features that appear similar in broadband visible light but differ in specific wavelengths. For instance, the Landsat program's Multispectral Scanner (MSS), operational since 1972, utilized four bands in the visible and near-: green (0.50–0.60 μm), red (0.60–0.70 μm), near- 1 (0.70–0.80 μm), and near- 2 (0.80–1.10 μm), all at approximately 60 m , facilitating detailed surface analysis over large areas. Later Landsat sensors, such as the Thematic Mapper, added bands including (0.45–0.52 μm) and shortwave up to 2.35 μm. Key applications of multispectral imaging include land cover classification, where bands in the visible and near-infrared spectrum help map urban expansion, forests, and water bodies by leveraging differences in albedo and vegetation reflectance. In agriculture, the Normalized Difference Vegetation Index (NDVI), calculated as NDVI = (NIR - Red) / (NIR + Red), quantifies vegetation health and biomass by exploiting the strong chlorophyll absorption in the red band (around 0.65 μm) and high reflectance in the near-infrared band (around 0.85 μm), with values typically ranging from -1 to 1 to indicate barren to dense vegetation. Mineral mapping benefits from shortwave infrared bands (e.g., 1.55-1.75 μm and 2.08-2.35 μm in Landsat), which reveal diagnostic absorption features of minerals like kaolinite and montmorillonite, aiding in resource exploration and environmental monitoring. Prominent sensors exemplify these capabilities: the (MODIS) on NASA's and Aqua satellites operates with 36 spectral bands from 0.4 to 14.4 μm, offering spatial resolutions from 250 m to 1 km, which supports global-scale applications like daily vegetation monitoring and . Advantages over panchromatic systems are evident in scenarios like discrimination, where shortwave infrared distinguishes clay-rich soils from sandy ones due to hydroxyl group absorptions around 2.2 μm, improving accuracy in thematic mapping.

Hyperspectral Applications

Hyperspectral imaging acquires data across more than 100 contiguous spectral bands, each typically narrower than 10 nm, spanning wavelengths from approximately 400 to 2500 nm to capture detailed spectral signatures of materials. This fine resolution enables the discrimination of subtle differences in reflectance, far surpassing multispectral approaches with coarser bands. A seminal example is the Airborne Visible/Infrared Imaging Spectrometer (AVIRIS), developed by NASA's Jet Propulsion Laboratory, which operates with 224 bands over 400–2500 nm for Earth remote sensing applications. Spaceborne hyperspectral sensors include the Italian Space Agency's PRISMA (launched 2019, 239 bands) and Germany's EnMAP (launched 2022, 242 bands), enabling global monitoring of environmental changes and resource mapping; recent commercial satellites like Planet's Tanager-1 (launched August 2024, over 400 bands) further enhance data accessibility as of 2025. Hyperspectral sensors commonly employ two scanning mechanisms: pushbroom systems, which image entire lines simultaneously as the platform moves forward, providing efficient coverage and stronger signal per pixel; and whiskbroom systems, which use a rotating mirror to scan points sequentially across a line, offering flexibility but potentially lower signal-to-noise ratios due to shorter dwell times. In , detects crop stress through minute changes in plant reflectance spectra, such as shifts in around 680 or indicators near 1450 , allowing early intervention to optimize yields and resource use. For , it identifies oil spills by matching unique spectral signatures in the shortwave infrared (e.g., features at 1720–1750 ), enabling rapid assessment of spill extent and from or platforms. In defense applications, hyperspectral data reveals camouflaged targets by exploiting mismatches between artificial materials and natural backgrounds across numerous bands, where multispectral methods might fail due to limited . Data processing in hyperspectral analysis often begins with spectral unmixing to decompose mixed pixels into constituent endmembers, based on the linear mixing model where the observed spectrum \mathbf{r} is approximated as \mathbf{r} = \sum_{i=1}^{p} f_i \mathbf{s}_i + \mathbf{n}, with f_i as abundance fractions (summing to 1), \mathbf{s}_i as endmember spectra, and \mathbf{n} as noise; this model assumes sub-pixel mixing without significant nonlinear interactions. techniques like () then address the high redundancy in hundreds of bands by projecting data onto fewer orthogonal components that retain most variance, facilitating classification and visualization without substantial information loss. Key challenges include managing massive data volumes, often reaching gigabytes for a single scene due to the hundreds of bands and high spatial resolutions, which demand efficient and storage strategies. Atmospheric correction is equally critical to remove and effects from gases like and aerosols; the MODTRAN radiative transfer model simulates these influences to convert at-sensor radiance to surface , though it requires accurate input parameters such as and altitude for reliable results.

Other Technical Applications

Audio Signal Processing

In audio signal processing, the audible frequency spectrum is typically divided into low-frequency and high-frequency bands to facilitate efficient , particularly at low bitrates where full-bandwidth encoding is impractical. For instance, the low band often covers frequencies up to approximately 8 kHz, while the high band extends from 8 kHz to 16 kHz or higher, depending on the application. Spectral band replication (SBR) addresses the limitations of coding by reconstructing the high-frequency content from the low-band signal using harmonic redundancy, thereby enhancing perceived audio quality without transmitting the full high-band spectrum. The SBR mechanism employs parametric coding, where the encoder analyzes the high-band and transmits a compact set of parameters—such as spectral shape, noise levels, and coupling information—alongside the core low-band data. At the , high frequencies are synthesized through techniques, such as shifting portions of the low-band upward by an or using patching to copy and adapt spectral segments, followed by application of the transmitted to match the original . This approach leverages the perceptual similarity between low and high harmonics in audio signals, enabling efficient reconstruction with minimal additional bitrate overhead of about 2-4 kb/s. SBR is integrated as an enhancement layer in codecs like (AAC) and High-Efficiency AAC (HE-AAC), where the core coder handles the low band and SBR operates post-. Applications of SBR are prominent in low-bitrate scenarios, such as streaming audio at 24 kb/s for speech enhancement or 64 kb/s for music, where it extends bandwidth to achieve near-full-range suitable for devices and . In , SBR improves intelligibility by restoring high-frequency consonants, while for music, it preserves richness in compressed streams. Perceptual is commonly evaluated using the (PEAQ) metric, standardized in BS.1387, which correlates scores with subjective listening tests and shows improvements for SBR-enhanced signals compared to non-extended audio at equivalent bitrates. SBR was developed in the late 1990s by Coding Technologies, a firm founded in 1997, as a novel extension method to boost compression efficiency in perceptual audio coders. It gained prominence through products like and was standardized in as part of MPEG-4 Audio Amendment 1, enabling its adoption in HE-AAC for applications like ().

Telecommunications and Radar

In telecommunications and radar systems, spectral bands refer to designated portions of the radio frequency (RF) spectrum allocated for specific uses to ensure efficient signal transmission, minimal interference, and reliable operation. The (ITU) classifies these bands using letter designations that correspond to frequency ranges, facilitating global standardization. For instance, the L-band spans 1 to 2 GHz and is commonly used for (GPS) signals due to its ability to penetrate foliage and provide long-range navigation accuracy. Similarly, the S-band covers 2 to 4 GHz and supports applications, where its moderate enables detection of and storm patterns over extended areas. The X-band, ranging from 8 to 12 GHz, is widely employed in for , offering high-resolution for precise aircraft tracking and collision avoidance in terminal areas. In higher frequencies, the Ka-band (26.5 to 40 GHz) plays a key role in wireless communications, enabling high-data-rate backhaul links and mobile services through its wide availability. These allocations prevent overlap, with assigned based on application needs—such as narrow bands for radar Doppler processing in the X-band to measure shifts in air traffic scenarios, or broader allocations in Ka-band to support the high throughput demands of modern wireless networks. A critical aspect of spectral band performance in these systems is the , which quantifies the received signal power to ensure reliable communication. The models this in free space as: P_r = P_t G_t G_r \left( \frac{\lambda}{4\pi d} \right)^2 where P_r is the received power, P_t is the transmitted power, G_t and G_r are the transmitter and receiver gains, \lambda is the , and d is the distance. This equation highlights how higher-frequency bands like Ka-band experience greater due to shorter wavelengths, necessitating higher gains or shorter ranges for viable links. Propagation characteristics vary significantly across bands, influencing system design. Lower bands like L- and S-band exhibit lower , supporting robust long-distance for GPS and weather monitoring even in adverse conditions. In contrast, millimeter-wave (mmWave) bands above 30 GHz, such as those explored for extensions beyond Ka-band, face challenges from increased atmospheric and , which can exceed 20 dB/km in rainy conditions, limiting range but enabling ultra-high data rates through massive bandwidths. These properties drive innovations like to mitigate attenuation in high-band and applications. Regulatory frameworks underpin spectral band management to avoid interference and promote equitable access. The U.S. Radio Act of 1927 established the (predecessor to the FCC) to allocate frequencies amid growing broadcast demands, marking the start of formalized spectrum governance. Today, the FCC and ITU oversee auctions for commercial bands, such as those in Ka-band for , generating billions in revenue while enforcing technical standards for power limits and emission masks. This international coordination ensures bands like S- and X-band remain dedicated to critical uses, balancing with public safety.

References

  1. [1]
    spectral band or spectral region - MicroImages
    spectral band or spectral region: A well-defined, continuous wavelength range in the spectrum of reflected or radiated electromagnetic energy.
  2. [2]
    Spectral Band - an overview | ScienceDirect Topics
    A spectral band is a specific range of wavelengths in the electromagnetic spectrum used for capturing and analyzing data.
  3. [3]
    Remote Sensing | NASA Earthdata
    Spectral resolution is the ability of an instrument to discern finer wavelengths, that is, having more and narrower bands. Many instruments are considered to be ...
  4. [4]
    Landsat 8 | U.S. Geological Survey - USGS.gov
    Two spectral bands: Band 10 TIRS 1 (10.6 - 11.19 µm) 100 m; Band 11 TIRS 2 (11.5 - 12.51 µm) 100 m. Landsat 8 Spacecraft Facts. Built by Orbital Science ...
  5. [5]
    What are the best Landsat spectral bands for use in my research?
    Landsat 8/9 Band 1 is for coastal/aerosol, Band 3 for vegetation, Band 5 for biomass, Band 6 for moisture, and Band 7 for improved moisture. Band 10 is for ...
  6. [6]
    [PDF] Spectroscopy and Remote Sensing - SPIE
    Spectroscopy and Remote Sensing. 7. Spectral bandwidth refers to the width of an individual spectral band—a wavelength or frequency-width—that can be ...
  7. [7]
    Types of Spectra and Spectroscopy - NASA Science
    Jul 11, 2018 · The dark lines and dim bands of light in a transmission spectrum correspond to atoms and molecules in the planet's atmosphere. The amount of ...
  8. [8]
    Chapter 6: Electromagnetics - NASA Science
    Jan 16, 2025 · Radio frequencies are divided into groups that have similar characteristics, called "bands," such as "S-band," "X-band," etc. The bands are ...
  9. [9]
    Spectral Band - an overview | ScienceDirect Topics
    Spectral bands refer to specific ranges of wavelengths in the electromagnetic spectrum that can be utilized in optical responses, particularly in ...
  10. [10]
    Band spectrum | McGraw Hill's AccessScience
    A spectrum consisting of groups or bands of closely spaced lines. Band spectra are characteristic of molecular gases or chemical compounds.
  11. [11]
    6.3: IR Spectrum and Characteristic Absorption Bands
    Dec 15, 2021 · The wavenumber is defined as the reciprocal of wavelength (Formula 6.3), and the wavenumbers of infrared radiation are normally in the range of ...
  12. [12]
    Electromagnetic Spectrum - Introduction - Imagine the Universe!
    The electromagnetic (EM) spectrum is the range of all types of EM radiation. Radiation is energy that travels and spreads out as it goes.
  13. [13]
    Atomic Absorption Spectroscopy vs Molecular ... - AZoOptics
    Jun 21, 2023 · Atomic absorption spectra consist of discrete absorption lines with dark spaces, in contrast to the continuous spectrum seen in white light ...
  14. [14]
    6.3: Line Spectra and the Bohr Model - Chemistry LibreTexts
    Jul 7, 2023 · Niels Bohr explained the line spectrum of the hydrogen atom by assuming that the electron moved in circular orbits and that orbits with only certain radii were ...
  15. [15]
    Oppenheimer's first paper: Molecular band spectra and a ...
    Mar 1, 2007 · Beginning early in the 20th century spectroscopists attributed the infrared band spectra emitted by diatomic molecules to quantum vibration and ...Missing: spectroscopy | Show results with:spectroscopy
  16. [16]
    Photosynthesis - Absorption Spectra - Chlorophyll a and b
    The absorption maxima of chlorophyll a are lambda= 430 and lambda= 662 nm, that of chlorophyll b are at 453 and 642 nm.Missing: peaks | Show results with:peaks
  17. [17]
    Absorption Spectrum of Chlorophyll
    The chlorophyll was extracted using 100% acetone and the absorption spectrum was determined over a range of 400-700 nm using a Spectronic 20D spectrophotometer.Missing: bands | Show results with:bands
  18. [18]
    Spectroscopy 101 – How Absorption and Emission Spectra Work
    The differences in spectra reflect the differences in the amount of energy that the atoms absorb or give off when their electrons move between energy levels.
  19. [19]
    Dynamic Stokes shift in green fluorescent protein variants - PMC - NIH
    The Stokes shift between the absorption and emission of a chromophore reflects the displacement in potential surface between the ground and excited states and ...
  20. [20]
    The Bouguer‐Beer‐Lambert Law: Shining Light on the Obscure - PMC
    The Beer‐Lambert law is unquestionably the most important law in optical spectroscopy and indispensable for the qualitative and quantitative interpretation ...
  21. [21]
    [PDF] Vibrational-Rotational Spectroscopy ( ) ( ) - MIT
    Absorption of mid-infrared light (~300-4000 cm-1):. • Molecules can change vibrational and rotational states. • Typically at room temperature, ...Missing: PQR | Show results with:PQR
  22. [22]
    [PDF] Spectroscopy 1: rotational and vibrational spectra
    The R branch consists of lines with ΔJ = +1: ˜ ν R. J( )= S v+1,J +1 ... Unlike in infrared spectroscopy, a Q branch is obtained for all linear molecules.
  23. [23]
    Atomic Spectroscopy - Spectral Line Shapes, etc. | NIST
    Oct 3, 2016 · Doppler broadening is due to the thermal motion of the emitting atoms or ions. For a Maxwellian velocity distribution, the line shape is ...Missing: Δν_D = (ν₀ / √( 2kT ln2 /
  24. [24]
    [PDF] 16 — Line Profiles I [Revision : 1.2]
    – To combine pressure/collisional broadening with natural broadening, simply add widths: ∆λp/c+nat. = ∆λp/c λ. +. ∆λnat λ. • Doppler broadening. – So far ...
  25. [25]
    [PDF] Saturated Absorption Spectroscopy - UF Physics
    the FWHM is called the natural linewidth and given the symbol Γ. Thus. Γ = γ. 2π. (2). Figure 1: The Lorentzian line shape profile for resonance absorption. For ...
  26. [26]
    [PDF] 16. Line shapes
    The Voigt Function. This is the general line shape describing the case when both Lorentzian and Gaussian broadening is present, e.g., pressure broadening and ...
  27. [27]
    [PDF] Lecture 6: Spectral Lineshapes
    The physical argument employed in establishing the Voigt profile is that the effects of Doppler & collision broadening are decoupled. ... FWHM of lineshape gives ...
  28. [28]
    [PDF] Absorption Features - NMSU Astronomy
    Thus, we see that the shape of the optical depth, Voigt profile behaves as a Gaussian function for small x (near the line core), and behaves as the wings of ...Missing: band | Show results with:band
  29. [29]
    [PDF] AFIT Scholar Collisional Broadening and Shift of Rubidium 4 F Line ...
    Several common forms of homogeneous broadening are natural broadening, pressure broadening, and Stark broadening. On the contrary, in an inhomogeneous ...<|separator|>
  30. [30]
    [PDF] Metal fluorescence lidar (light detection and ranging) and the middle ...
    Homogeneous broadening refers to a mechanism that broadens the individual spectral lines of each atom in the same manner. An atomic transition's natural ...
  31. [31]
    Intro. to Signal Processing:Deconvolution - University of Maryland
    ... least-squares curve fitting of a proposed peak model to the data set. However, that process is actually conceptually distinct from Fourier deconvolution ...Missing: bands | Show results with:bands
  32. [32]
    An Improved LSD Technique for Accurate Line Profile Retrieval. - arXiv
    Mar 7, 2024 · Least-squares deconvolution (LSD) offers an alternative technique that deconvolves the entire spectrum to produce a single mean-weighted line ...
  33. [33]
    Mathematical Deconvolution of Mineral Absorption Bands - NASA ADS
    ... spectra using a least-squares fitting procedure. Gaussian Model. Early work tested the practicality of using Gaussian distributions, as was done by many ...
  34. [34]
    [PDF] Vibrational Bands of H, Ny Oz Molecules - Standard Reference Data
    The basic role of infrared spectroscopy in probing the. H, N, O, molecules is well established, in particular in the atmospheric area. As a consequence of the ...<|control11|><|separator|>
  35. [35]
    First Laboratory Detection of Vibration-Rotation Transitions of ... - NIH
    Interestingly, in our IR scans we also observed rovibrational transitions of the complex CH+-He (and similarly 13CH+-He) which are detected by their ...
  36. [36]
    The Nobel Prize in Physics 1975 - NobelPrize.org
    The Nobel Prize in Physics 1975 was awarded jointly to Aage Niels Bohr, Ben Roy Mottelson and Leo James Rainwater for the discovery of the connection between ...
  37. [37]
    Nuclear deformation and rotations - Book chapter - IOPscience
    The K quantum number is used as the key quantum number for discussing rotational bands. K isomerism is sketched. Moments of inertia and a caution about the ...
  38. [38]
    [PDF] arXiv:2409.02279v1 [nucl-th] 3 Sep 2024
    Sep 3, 2024 · Applying angular-momentum pro- jection yields a rotational band with a head that has the nuclear spin/parity Iπ = Kπ. As the spherical shell ...
  39. [39]
    Energy differences of ground state and γ 1 bands as a hallmark of ...
    The systematics of experimental energy differences between the levels of the ground state band and the even levels of the band in even-even nuclei are studied
  40. [40]
    Rotational bands in the quadrupole-octupole collective model
    Feb 16, 2018 · A collective band of positive as well as negative parity could be composed of vibrational and rotational motions. The octupole vibrational ...
  41. [41]
    [PDF] Lecture 4: Nuclear Structure 2 - INPP - Ohio University
    Nuclear matter is nearly incompressible, so the monopole oscillation takes a good bit of energy to excite. For even-even nuclei, the monopole oscillation ...
  42. [42]
    Full article: Fifty Years of Backbending - Taylor & Francis Online
    Jun 14, 2022 · The revolutionary spectrum of the rotational band of the nucleus 160Dy was published 51 years ago [Citation8] (see. Figure 2 ). In the paper, it ...Missing: variations | Show results with:variations
  43. [43]
    [PDF] arXiv:2003.07902v1 [nucl-th] 17 Mar 2020
    Mar 17, 2020 · High-spin rotational bands in rare-earth Er (Z = 68), Tm (Z = 69) and Yb (Z = 70) iso- topes are investigated by three different nuclear models.
  44. [44]
    Coulomb excitation and transfer reactions with rare neutron-rich ...
    Neutron-rich radioactive ion beams available from the HRIBF allow a variety of measurements around the 132Sn region, including Coulomb excitation, ...Missing: bands | Show results with:bands
  45. [45]
    [PDF] Coulomb excitation of exotic nuclei at REX-ISOLDE with MINIBALL
    Most important method is the γ-ray spectroscopy following “safe” Coulomb excitation and, more recently, nucleon-transfer reactions. The experimental setup ...
  46. [46]
    Hyperspectral Imaging - an overview | ScienceDirect Topics
    Hyperspectral sensors acquire more than 100 contiguous spectral bands with a narrow bandwidth (5–10 nm) in a wavelength region between 500 and 2500 nm.
  47. [47]
    AVIRIS - Airborne Visible / Infrared Imaging Spectrometer
    May 22, 2024 · ... spectral channels (also called bands) with wavelengths from 400 to 2500 nanometers (nm). AVIRIS has been flown on four aircraft platforms ...
  48. [48]
  49. [49]
    A systematic review of hyperspectral imaging in precision agriculture
    Hyperspectral imaging (HSI) is data rich in solving agricultural problems like disease detection, weed detection, stress detection, crop monitoring, nutrient ...
  50. [50]
    Hyperspectral Imaging-Based Marine Oil Spills Remote Sensing ...
    The detection system developed in this study provides a robust means for oil spill monitoring on drilling platforms in remote sensing of the marine environment.
  51. [51]
    Application of Hyperspectral Imager in Disguised Target ... - Optosky
    Jan 10, 2023 · According to this difference, hyperspectral remote sensing may detect camouflaged targets. The role of hyperspectral imagers in national ...
  52. [52]
    Hyper-spectral Unmixing algorithms for remote compositional ...
    Oct 13, 2025 · The linear mixing model (LMM) approximates each pixel spectrum y j with a linear combination of end member spectra, with the corresponding ...
  53. [53]
    Dimensionality Reduction and Classification of Hyperspectral ...
    PCA is a linear transformation unsupervised dimensionality reduction algorithm. It maintains data information and simplifies the dimension of hyperspectral ...
  54. [54]
    [PDF] Status of Atmospheric Correction Using a Modtran4-based Algorithm
    Atmospheric correction (or compensation) of AVIRIS or other spectral imagery refers to the retrieval of surface reflectance spectra from measured radiances.
  55. [55]
    [PDF] SBR explained: White paper
    SBR (Spectral Band Replication) is a new audio coding enhancement tool. It offers the possibility to improve the performance of low bitrate audio and speech ...
  56. [56]
    Spectral Band Replication - MultimediaWiki - Multimedia.cx
    Mar 26, 2011 · Spectral Band Replication or SBR is a lossy scheme developed by Coding Technologies (now part of Dolby) to efficiently reconstruct the high ...Missing: invented | Show results with:invented
  57. [57]
    An Overview of the Coding Standard MPEG-4 Audio Amendments 1 ...
    These amendments concern parametric coding techniques and encompass Spectral Band Replication (SBR), Sinusoidal Coding (SSC), and Parametric Stereo (PS).Missing: history | Show results with:history
  58. [58]
    [PDF] Perceived Audio Quality Analysis in Digital Audio Broadcasting Plus ...
    One of widely used objective methods is Perceptual. Evaluation of Audio Quality (PEAQ). PEAQ is defined in Recommendation ITU-R BS.1387 [19]. It uses ear ...
  59. [59]
    All Bands of Satellite Frequencies and their Applications | Journal
    L-band frequencies range from 1-2 GHz and are chiefly used for GPS navigation and connecting satellite phones. They have longer wavelengths and are good at ...
  60. [60]
    What are S- and X-Band Radar and How Are They Used Today?
    Dec 19, 2022 · S-band radar is better for special applications and long-range detection. It is used more for surveillance as it provides better coverage.
  61. [61]
    All About the 5G Ka-Band for Mobile Services
    Mar 7, 2023 · The 5G low-band spectrum is employed for traditional coverage applications and specific use cases such as industrial automation, the Internet of ...
  62. [62]
    On challenges of sixth-generation (6G) wireless networks
    5G bands have several challenges including poor foliage penetration, atmospheric and free space path loss, and high deployment cost. Table 3 shows a ...
  63. [63]
    Radio Act of 1927 [established the Federal Radio Commission]
    The Radio Act of 1927, Pub. Law 632, 69th Concress, February 23, 1927, established the Federal Radio Commission, predecessor to the FCC in 1934.
  64. [64]
    The Federal Communications Commission's Spectrum Auction ...
    In 1993, Congress authorized the FCC to use competitive bidding (ie, auctions) to grant licenses for rights to use specific frequencies for commercial wireless ...