Fact-checked by Grok 2 weeks ago

Absorption spectroscopy

Absorption spectroscopy is a spectroscopic technique that measures the absorption of by atoms, molecules, or other materials as a function of or , producing an absorption spectrum that reveals characteristic patterns akin to fingerprints for identifying and quantifying substances in a sample. The fundamental principle underlying absorption spectroscopy is the interaction of photons with matter, where radiation is absorbed when its matches the difference between quantum levels in the sample, leading to transitions such as electronic, vibrational, or rotational excitations. This process is quantitatively described by Beer's Law, which states that the (A) of is directly proportional to the concentration (c) of the absorbing species, the path length (d) through the sample, and the molar absorptivity (ε), expressed as A = εcd. Historically, the technique traces its roots to the early with von Fraunhofer's observations of dark lines in the solar , laying the groundwork for modern applications. Absorption spectroscopy encompasses various types, including ultraviolet-visible (UV-Vis) for electronic transitions, infrared (IR) for vibrational modes, and X-ray for core electron excitations, each tailored to specific wavelength ranges and sample properties. Its applications span for qualitative identification and of compounds in solutions, for detecting pollutants, astronomical studies to analyze stellar compositions, and biomedical research for probing molecular structures. Widely used due to its sensitivity and versatility, the method is essential in fields like pharmaceuticals, , and , enabling non-destructive characterization of samples at trace levels.

Principles

Basic Mechanism of Absorption

Absorption spectroscopy is based on the process by which atoms or molecules absorb photons of , resulting in the excitation of electrons, vibrations, or rotations from lower to higher energy states. This phenomenon was first systematically observed in 1814 by , who identified hundreds of dark absorption lines in the solar spectrum using a , laying the groundwork for understanding spectral absorption in stellar atmospheres. At the quantum mechanical level, absorption occurs through discrete transitions between quantized energy levels, where the photon's must match the difference between initial and final states, given by E = h\nu, with h as Planck's and \nu as frequency. These transitions are governed by selection rules derived from , particularly for electric dipole interactions, which require a change in orbital angular momentum quantum number of \Delta l = \pm 1, such as s to p transitions in atoms. The probability of absorption at a given wavelength depends on the gap between levels, with shorter wavelengths corresponding to larger energy differences typically involving excitations. Molar absorptivity, denoted \epsilon, quantifies the intrinsic probability of these transitions for a specific or at a given , reflecting how strongly the interacts with the . Common examples include in the ultraviolet-visible (UV-Vis) region, where valence electrons are promoted (e.g., \pi \to \pi^* in molecules), and vibrational in the () region, involving changes in molecular bond stretching or modes. The Beer-Lambert law provides a quantitative framework for relating this intensity to concentration, but the underlying mechanism remains rooted in these quantum transitions.

Beer-Lambert Law

The Beer-Lambert law, a foundational principle in absorption spectroscopy, combines the contributions of and August Beer in quantifying absorption by matter. In 1760, demonstrated that the intensity of diminishes exponentially as it passes through a homogeneous medium, with the decrease proportional to the distance traveled. August Beer extended this in 1852 by applying it to colored solutions, showing that absorption also depends linearly on the concentration of the absorbing solute, thus enabling in solution-based measurements. The law states that absorbance A is given by A = \epsilon c l, where \epsilon is the absorptivity (specific to the substance and ), c is the of the absorber, and l is the . This form arises from the describing infinitesimal intensity loss, dI = -\alpha I \, dl, where \alpha = \epsilon c is the linear absorption coefficient; integrating from incident intensity I_0 to transmitted intensity I over path length l yields the exponential relation I = I_0 e^{-\epsilon c l}. Absorbance links to T = I / I_0 via the base-10 logarithmic form A = -\log_{10} T, which linearizes the for practical plotting and analysis, ensuring A increases linearly with c or l under ideal conditions. The molar absorptivity \epsilon typically carries units of L ⁻¹ cm⁻¹, with c in L⁻¹ and l in cm, rendering A dimensionless and facilitating calibration curves for concentration determination in analytical applications. The law relies on key assumptions, including monochromatic incident light, dilute solutions to avoid solute interactions, and negligible , , or reflection within the sample. Limitations arise when these are violated, such as in polychromatic illumination causing nonlinear deviations, or high concentrations inducing inner effects where emitted or scattered light is reabsorbed, distorting measured . Additionally, chemical equilibria in the sample can shift effective c, further compromising beyond dilute regimes.

Absorption Spectra

Characteristics and Interpretation

An absorption spectrum is a graphical representation of the absorbance of light by a sample as a function of wavelength or frequency, typically plotted with absorbance on the y-axis and wavelength (in nanometers) or wavenumber (in cm⁻¹) on the x-axis. This plot reveals the specific wavelengths where the sample absorbs energy, corresponding to transitions between molecular energy levels, such as electronic excitations in ultraviolet-visible (UV-Vis) spectroscopy or vibrational modes in infrared (IR) spectroscopy. Absorbance values generally range from 0 (no absorption) to 2 (near-complete absorption at that wavelength), providing a quantitative measure of light interaction with the analyte. Key features of absorption spectra include prominent peaks, known as absorption bands, which indicate the wavelengths of maximum due to specific molecular transitions. These bands arise from , vibrational, or rotational excitations, with their positions serving as unique signatures for molecular structures. Bandwidths, or the width of these peaks, reflect the of the transitions and often incorporate from vibrational or rotational sublevels, though they can broaden due to environmental influences. Baseline corrections are essential to establish a reference level for zero , subtracting any background signal from the or sample matrix to ensure accurate peak identification. Interpretation of absorption spectra relies on analyzing peak positions to identify functional groups and integrated peak intensities to quantify concentrations. For instance, in IR spectroscopy, a strong absorption band around 1700 cm⁻¹ typically indicates a carbonyl (C=O) stretch, allowing deduction of molecular composition through comparison with known spectral libraries. Peak positions provide qualitative insights into bonding environments, while the area under the curve or peak height, when combined with path length and molar absorptivity, enables concentration determination via the Beer-Lambert law. This approach prioritizes conceptual matching of spectral patterns to molecular motifs rather than exhaustive cataloging. Several factors influence the appearance of absorption spectra, necessitating controlled conditions for reliable interpretation. Solvent effects can shift peak positions (bathochromic or hypsochromic shifts) and alter bandwidths by interacting with the solute's electronic or vibrational states, such as through or hydrogen bonding. Temperature broadening occurs as increased populates higher vibrational levels, leading to wider peaks and reduced of fine structure. Additionally, pH variations can modify states of ionizable groups, thereby changing absorption characteristics, particularly in UV-Vis spectra of acids or bases. Absorption spectra are often presented in normalized forms, such as molar absorptivity (ε) versus , to facilitate comparison across samples by removing concentration dependencies. Difference spectra, obtained by subtracting a from the sample, highlight changes over time or conditions, aiding kinetic studies of reactions without baseline drift issues. These methods enhance and emphasize subtle variations for precise molecular .

Relation to Other Spectra

Absorption spectra are closely related to transmission spectra, as both describe the interaction of with a sample but from complementary perspectives. The T, defined as the ratio of transmitted to incident (T = I/I_0), is directly related to A through the Beer-Lambert law, where T = 10^{-A}. This inverse relationship means that regions of high correspond to low , and the two spectra are mathematically equivalent in many contexts, though transmission spectra emphasize the that passes through the sample while focuses on the removed. In practice, the terms are sometimes used interchangeably for transparent media, but measurements prioritize quantifying the absorbing species, whereas highlights overall . Absorption spectra complement emission spectra, which arise from the relaxation of back to the . Absorption promotes molecules from the to by absorbing photons, while involves the subsequent release of as , often resulting in a red-shifted relative to absorption due to vibrational relaxation in the , known as the . This shift, typically measured as the difference between the absorption and maxima, ensures that wavelengths are longer than those absorbed, preventing overlap and enabling phenomena like . Absorption spectra can predict possible lines by identifying allowed electronic transitions, providing a foundational understanding of the levels involved in both processes. Scattering spectra, such as those from Raman spectroscopy, probe similar molecular vibrations as infrared (IR) absorption but through different mechanisms. In IR absorption, vibrations are excited directly by light if they induce a change in the molecular dipole moment, limiting observations to polar bonds. Raman scattering, however, involves inelastic scattering of light where the incident photon exchanges energy with the molecule, requiring a change in polarizability rather than dipole moment, thus allowing detection of non-polar bonds that are IR-inactive. This complementarity enables Raman to access vibrational information akin to IR absorption without the selection rule restrictions, though both techniques map out molecular energy levels in the vibrational regime. Reflection spectra are linked to through surface and interactions, particularly in opaque or media like solids. contributes to by attenuating that penetrates the , and the Kubelka-Munk theory relates the diffuse R to the coefficient K and coefficient S via the function F(R) = (1 - R)^2 / (2R) \approx K/S, approximating from surface-reflected . However, while spectroscopy measures properties throughout the sample depth, spectra primarily capture interfacial effects and at the surface, making the techniques distinct for heterogeneous s. In practical applications, absorption spectroscopy excels in quantitative bulk analysis of concentrations via the Beer-Lambert law, offering high sensitivity for trace analytes in solutions or gases. Emission spectroscopy, by contrast, is often favored for qualitative identification of elements or species due to its bright, line-specific signals from excited atoms or molecules, though it may require higher sample energies. These distinctions guide the choice of technique based on whether precise quantification or spectral fingerprinting is prioritized.

Experimental Techniques

Instrumentation and Setup

Absorption spectroscopy instrumentation typically consists of several core components that facilitate the measurement of absorption by a sample across specific ranges. The source provides the incident ; for ultraviolet-visible (UV-Vis) spectroscopy, common sources include arc lamps for the UV region (approximately 190–350 nm) and tungsten-halogen lamps for the visible region (approximately 350–900 nm). In () spectroscopy, thermal sources such as globars or Nernst glowers emit broadband from about 400 nm to 50 μm. A or interferometer then selects or disperses wavelengths: in dispersive UV-Vis systems, holographic gratings or prisms isolate narrow bands, while IR setups often employ interferometers to enable simultaneous detection of all wavelengths via processing. The sample is held in a or cell for liquids, with fused silica preferred for UV transparency down to 190 nm and or for visible ranges; path lengths are typically 1–10 cm to optimize signal according to the Beer-Lambert law, which underpins calculations. Detectors convert transmitted to electrical signals, such as tubes for high-sensitivity UV-Vis detection or thermal detectors like pyroelectric devices for IR. Instrument configurations vary between single-beam and double-beam spectrophotometers to address practical challenges like source instability. In a single-beam setup, the passes sequentially through the sample and , offering simplicity, lower cost, and higher light throughput for sensitive measurements, though it requires manual switching and is susceptible to drift. Double-beam systems split the beam using a or , directing one path through the sample and the other through a , which automatically compensates for fluctuations in source intensity or detector response, improving accuracy over long scans. Modern instruments incorporate digital controls for automated selection, , and beam alignment, enhancing . Sample preparation ensures uniform interaction with the light beam and minimizes or . For liquid solutions, samples are diluted in solvents compatible with the wavelength range (e.g., avoiding UV-absorbing solvents like below 250 nm) and placed in matched cuvettes to control path length precisely. Gaseous samples require sealed cells with windows transparent to the spectral region, often with adjustable path lengths up to several meters for low-concentration detection. Solids are prepared as dispersions, such as (KBr) pellets for mid- analysis, where the sample (1–2% by weight) is ground finely with KBr and pressed under vacuum to form a transparent disk approximately 1 mm thick. Alternatively, (ATR) is widely used for solids and liquids in spectroscopy, involving direct contact with a (e.g., or ZnSe) to measure surface without extensive preparation, reducing issues like moisture sensitivity in KBr methods. Path length control is critical in all cases to maintain linearity in readings. The basic experimental setup involves initial alignment of optical components to maximize light throughput, followed by recording the incident intensity I_0 using a blank (solvent or empty cell) across the desired wavelength range. The sample is then inserted, and transmitted intensity I is measured by scanning wavelengths, typically from low to high to avoid heating effects in IR. Data processing computes absorbance as A = \log_{10}(I_0 / I), directly relating to concentration via the Beer-Lambert law. Safety protocols emphasize protection from intense light sources, such as UV-blocking eyewear for lamps to prevent eye damage, and proper for volatile solvents or sources that generate heat. Calibration ensures reliability: accuracy is verified using standards like oxide solution, which exhibits sharp peaks at known positions (e.g., 241.5 , 287.2 ), while is minimized through clean and enclosure design to keep levels below 0.1% across the spectrum. Photometric linearity is checked with neutral density filters or solutions, and modern digital interfaces allow real-time monitoring and automated corrections for baseline drift.

Specific Methods

Ultraviolet-visible (UV-Vis) absorption spectroscopy operates in the wavelength range of approximately 200-800 nm, where absorption arises from electronic transitions in molecules, such as π → π* and n → π* excitations in organic compounds. This method commonly employs diode array detectors to enable rapid scanning and simultaneous acquisition of full spectra across the range. Infrared (IR) absorption spectroscopy targets the mid-IR region from 4000 to 400 cm⁻¹, corresponding to vibrational modes of molecular bonds, including and transitions that provide fingerprints for functional groups. Traditional dispersive IR uses a or to separate wavelengths, while IR (FTIR) employs a with a moving mirror and beam splitter to generate an interferogram, which is Fourier-transformed into a for higher resolution and speed. FTIR offers advantages in and multiplex detection (Felgett advantage), but IR methods generally suffer from strong interference due to O-H bands overlapping with sample signals. Atomic absorption spectroscopy (AAS) focuses on by atomizing samples in a or to produce ground-state atoms, which absorb light from element-specific hollow cathode lamps emitting narrow atomic lines. In AAS, aspiration into an air-acetylene achieves rapid atomization for high-throughput measurements, whereas AAS provides lower detection limits through controlled heating stages for trace elements. Near-infrared (NIR) absorption spectroscopy covers 700-2500 nm, probing and bands of vibrations, which are weaker but enable non-destructive analysis of solids, liquids, and tissues without extensive preparation. Emerging methods include terahertz absorption spectroscopy, which examines low-energy vibrational and rotational modes in the 0.1-10 THz range for applications in biomolecules and materials. Time-resolved absorption spectroscopy uses - configurations to study reaction kinetics by exciting samples with a pulse and monitoring transient absorption with a delayed .

Applications

Chemical Analysis

Absorption spectroscopy plays a central role in quantitative chemical analysis by enabling precise determination of concentrations through adherence to the Beer-Lambert law, which states that is directly proportional to concentration under conditions of constant path length and absorptivity. In pharmaceutical applications, this principle is routinely applied to assess drug purity; for instance, UV-Vis measures the of active pharmaceutical ingredients at characteristic wavelengths, allowing calculation of concentrations. The limit of detection (LOD) in such analyses is governed by the , typically defined as three times the standard deviation of the blank divided by the slope of the , for photometric methods in routine laboratory settings. For qualitative identification, absorption spectra act as molecular fingerprints, facilitating compound verification by comparing observed λ_max and spectral shapes to . In UV-Vis spectroscopy, this is particularly effective for conjugated systems, where the extent of π-conjugation shifts absorption maxima to longer wavelengths (bathochromic shift), enabling distinction of structural isomers or functional groups without prior separation. Such spectral matching confirms identity in or impurity profiling, with high specificity derived from the unique transitions involved. Absorption spectroscopy also supports the study of reaction kinetics and mechanisms by monitoring temporal changes in , which correlate directly with reactant or product concentrations via Beer's law. This time-resolved approach yields rate constants and reaction orders; for example, in homogeneous reactions, absorbance at a fixed tracks species evolution, revealing pseudo-first-order kinetics with rate constants determined from fits. Practical examples include spectrophotometric , where continuous absorbance measurements during reagent addition identify equivalence points through curve inflections, applicable to systems where traditional indicators fail, such as titrations of absorbing analytes or products alone. In complexometric analysis, metallochromic indicators like form colored complexes with metal ions, and absorbance shifts upon titration with EDTA signal endpoint detection, enhancing selectivity for divalent cations in aqueous solutions. Modern advances in chemical analysis have integrated absorption spectroscopy with chromatographic separations in hyphenated techniques, such as coupled to UV detection (HPLC-UV), which resolves complex mixtures before spectral quantification. These developments have enhanced and reduced analysis times for pharmaceutical formulations, enabling simultaneous and purity assessment in multi-component samples.

Astronomy and Remote Sensing

Absorption spectroscopy plays a pivotal role in astronomy by enabling the analysis of stellar atmospheres through the observation of , which are dark absorption features in the continuous spectrum of stars caused by cooler gases absorbing specific wavelengths. These lines, first systematically cataloged by in 1814, reveal the , , and of stellar atmospheres by identifying and molecular species such as , , and metals. For instance, the in the solar spectrum allow astronomers to determine the abundance of elements like iron and calcium, providing insights into the star's physical conditions. In the (), absorption spectroscopy detects molecular (H₂) through its absorption lines, which trace the distribution and properties of diffuse gas clouds along lines of sight to background stars. These H₂ lines, observed primarily in the far-UV range (around 912–1108 Å), indicate the presence of cold, dense regions where molecules form, helping to map the ISM's structure and dynamics. Additionally, during exoplanet transits, when a passes in front of its host star, absorption spectroscopy measures the planet's atmospheric transmission spectrum, revealing molecular signatures like water vapor, , and through depth variations in the stellar light. This technique has identified atmospheric compositions in hot Jupiters, such as enhanced water absorption in . Remote sensing applications extend absorption spectroscopy to Earth's atmosphere, where techniques like measure trace gases such as by analyzing narrow-band UV absorptions in scattered sunlight or direct beams over long paths. DOAS exploits the distinct absorption fingerprints of in the 280–340 nm range to quantify column densities, aiding in monitoring stratospheric depletion. Satellite-based instruments, including the Ozone Monitoring Instrument (OMI) launched in 2004 aboard NASA's satellite, use nadir-viewing UV spectroscopy to map global distributions with resolutions of 13×24 km, supporting climate and air quality assessments. Key challenges in these applications include in stellar atmospheres, arising from thermal motions of atoms that widen lines and complicate abundance determinations, requiring high spectral resolutions exceeding 10,000 to resolve profiles. In atmospheric , long path lengths (up to thousands of kilometers for satellite views) introduce multiple scattering and variable optical depths, necessitating corrections for interference and photon path distributions to achieve accurate retrievals. often relies on measurements, defined as the integrated area of an line normalized by the flux, to quantify line strengths and derive abundances via curve-of-growth methods, as applied in high-resolution stellar spectra. Seminal discoveries underscore the technique's impact: in 1868, and independently detected helium's yellow emission line (587.6 nm) in the Sun's during a total , marking the first identification of an unknown on . More recently, since 2022, the (JWST) has advanced molecular mapping through mid-infrared transmission spectroscopy, detecting and in exoplanet atmospheres like WASP-39b, revealing diverse chemical environments at unprecedented sensitivity. As of 2025, JWST observations have provided new constraints on potential biosignatures, such as dimethyl sulfide (DMS) in the atmosphere of K2-18 b.

Biological and Medical Uses

Absorption spectroscopy plays a central role in quantifying biomolecules, particularly proteins and nucleic acids, through their characteristic ultraviolet (UV) absorbance. Proteins exhibit strong absorption at 280 nm primarily due to aromatic amino acids such as tryptophan and tyrosine, enabling rapid and non-destructive concentration measurements in biological samples. Similarly, DNA and RNA absorb maximally at 260 nm owing to the π-π* transitions in their nucleotide bases, allowing for straightforward nucleic acid quantification without the need for dyes or labels. The ratio of absorbance at 260 nm to 280 nm (A260/A280) serves as a purity indicator, with values around 1.8 for DNA and 2.0 for RNA signifying minimal protein contamination. In medical diagnostics, absorption spectroscopy underpins non-invasive techniques for monitoring vital parameters. employs and red (typically 660 nm and 940 nm) to differentiate oxygenated and deoxygenated based on their distinct absorption spectra, providing real-time peripheral (SpO2) measurements essential for respiratory assessment. For diagnosis, particularly in neonates, transcutaneous bilirubinometry uses visible absorption around 450-460 nm, where exhibits peak absorbance, to estimate serum levels and guide phototherapy decisions without invasive blood draws. These methods highlight absorption spectroscopy's utility in point-of-care settings, reducing patient discomfort while maintaining clinical accuracy. Advanced applications extend to imaging and therapeutic monitoring, leveraging absorption contrasts . Photoacoustic imaging combines optical absorption with ultrasonic detection, where NIR laser pulses induce thermoelastic expansion in absorbing tissues (e.g., hemoglobin-rich vasculature), generating acoustic waves for high-resolution anatomical and functional imaging in and . In pharmacokinetics, UV-Vis absorption tracks drug metabolites by monitoring changes in absorbance spectra during and assays, aiding in the assessment of metabolic stability and . Enzyme activity assays, such as those measuring NADH production at 340 nm in dehydrogenase reactions, exemplify real-time kinetic monitoring critical for and biochemical research. Recent advancements in the 2020s focus on portable devices for non-invasive glucose monitoring, addressing challenges. These systems analyze absorptions in the 1000-1700 nm range from glucose's C-H and O-H bonds in interstitial fluid, with algorithms improving prediction accuracy despite inter-individual variability. Clinical implementation requires rigorous calibration to account for biological factors like pigmentation, , and motion artifacts, ensuring equitable performance across diverse populations and minimizing diagnostic biases. absorption methods briefly reference vibrational modes for biomolecular , complementing UV-Vis in comprehensive profiling.

References

  1. [1]
    Spectroscopy: A Measurement Powerhouse | NIST
    Mar 28, 2025 · In absorption spectroscopy, scientists measure light that has passed through clouds of atoms or molecules. Each type of atom or molecule ...
  2. [2]
    [PDF] A Brief Review on Absorption Spectroscopy - Hilaris Publisher
    Nov 26, 2021 · Absorption spectroscopy is additionally utilized in investigations of sub-atomic and nuclear physical science, cosmic spectroscopy and remote ...
  3. [3]
    [PDF] Absorption Spectrophotometry
    Absorption spectroscopy is a means of investigating samples by the energy which is absorbed in raising its molecules to excited states. While the phenomenon.
  4. [4]
    [PDF] Absorption spectroscopy at the limit: detection of a single atom
    Absorption spectroscopy dates back to the early 19th century when Fraunhofer observed characteristic lines in the solar spectrum (Fraunhofer lines) and to.
  5. [5]
    Absorption spectroscopy - PubMed
    Absorption spectroscopy is one of the most widely used techniques employed for determining the concentrations of absorbing species (chromophores) in solutions.Missing: principles | Show results with:principles
  6. [6]
    [PDF] Chapter 10 - DePauw University
    Another important application of molecular absorption spectroscopy is the determination of equilibrium constants. Let's consider, as a simple example, an ...
  7. [7]
    [PDF] in 1814, Joseph von Fraunhofer discovered dark lines in the solar ...
    In 1814, Fraunhofer used a glass prism to split sunlight into its rainbow colors. To his amazement, the resulting color fan contained about six hundred dark ...
  8. [8]
    [PDF] Module 1: Fundamentals of Spectroscopy - MIT OpenCourseWare
    Quantum mechanical particle-in-a-box model of electronic excited state energies. We've seen how absorption spectra have peaks at the resonance frequencies ω0 of.<|control11|><|separator|>
  9. [9]
    [PDF] Synchrotron X-ray Absorption Spectroscopy
    Selection Rules for X-ray absorption near-edge spectra. 2p →nf. 1s →nd. Weak. Δl=±2. Quadrupole. 2p →nd. 1s →np. Intense. Δl=±1. Dipole. LIII-edge. K-edge.
  10. [10]
    The Bouguer‐Beer‐Lambert Law: Shining Light on the Obscure - PMC
    Historical Overview. The work of Pierre Bouguer and Johann Heinrich Lambert dealing with spectrophotometry of the atmosphere was well known to August Beer as ...
  11. [11]
    Beer's Law - Theoretical Principles
    Where A is absorbance (no units, since A = log10 P0 / P ) e is the molar absorbtivity with units of L mol-1 cm-1 b is the path length of the sample - that ...
  12. [12]
    Inner filter effect in fluorescence spectroscopy: As a problem and as ...
    According to Lambert-Beer's law, light radiation drops exponentially inside the fluorescence cell [[4], [5], [6], [7]]. The expression for intensity of ...Missing: limitations | Show results with:limitations
  13. [13]
    UV-Visible Spectroscopy - MSU chemistry
    Absorbance usually ranges from 0 (no absorption) to 2 (99% absorption), and is precisely defined in context with spectrometer operation.
  14. [14]
    IR Spectrum and Characteristic Absorption Bands – Organic Chemistry
    IR spectra show absorption bands, with different functional groups having characteristic frequencies. The fingerprint region below 1500 cm-1 is unique to each ...
  15. [15]
    Spectroscopy 101 – How Absorption and Emission Spectra Work
    In the visible part of the spectrum, hydrogen absorbs light with wavelengths of 410 nm (violet), 434 nm (blue), 486 nm (blue-green), and 656 nm (red). Each of ...
  16. [16]
    [PDF] Fundamentals of modern UV-visible spectroscopy - Service Catalog
    Other parameters, such as pH and temperature, also may cause changes in both the intensity and the wavelength of the absorbance maxima. Conjugating the double ...
  17. [17]
    [PDF] Temperature-dependent vibrational dephasing - Stanford University
    Aug 8, 2002 · Frequency-selected vibrational echo experiments were used to investigate the temperature dependences of vibrational dephasing associated ...
  18. [18]
    Beer–Lambert law for optical tissue diagnostics: current state ... - NIH
    Oct 28, 2021 · Historical development of updating this law by taking into account specific additional factors such as light scattering and photon ...
  19. [19]
    [PDF] Atmospheric Transmission Beer's Law
    Mar 31, 2010 · Hence Beer's Law states that there is a linear relationship between absorbance and concentration (as opposed to a logarithmic dependence on ...
  20. [20]
    Relations between absorption, emission, and excited state chemical ...
    May 28, 2021 · In condensed phases, thermalization shifts the emission spectrum to lower frequency than the absorption spectrum, this is known as Stokes' ...
  21. [21]
    [PDF] Comparison between IR and Raman
    IR Selection Rule: For any IR absorption, the vibrational excitation must change the dipole moment of the molecule in order for radiation absorption to occur.
  22. [22]
    Complementary vibrational spectroscopy - PMC - NIH
    IR absorption spectroscopy, which is active for polar bonds such as O–H or N–H, is often used for identification of functional groups of molecules, while Raman ...
  23. [23]
    The Interpretation of Diffuse Reflectance Spectra - PMC - NIH
    We have seen that the Kubelka-Munk theory leads to a solution of the form, ... absorption coefficient from reflectance measurements by connecting the Kubelka-Munk ...
  24. [24]
    [PDF] Absorption, Emission and Fluorescence Spectroscopies
    Absorption is the process that consumes a photon and puts the atom or molecule in an excited state. Emission is the process that creates a photon and takes the ...Missing: practical | Show results with:practical
  25. [25]
    [PDF] UV-Vis Instrumentation
    Detectors for UV-VIS​​ Photon Transducers: Covert photon energy to electrical signal (current, voltage, etc.)
  26. [26]
    Other Types of Spectrophotometers
    In addition to a single beam spectrophotometer, which has only one beam of light, a double beam spectrophotometer may also be used. A double beam ...
  27. [27]
    [PDF] Method 7000B: Flame Atomic Absorption Spectrophotometry, part of ...
    This principle is the basis of atomic absorption spectrophotometry. 3.0 DEFINITIONS. Refer to Chapter One, Chapter Three, and the manufacturer's instructions ...
  28. [28]
    Vibrational Spectroscopy Chapter 4 - Forest Products Laboratory
    ... infrared (NIR) spectroscopy ... and that result in characteristic' absorption (infrared) [4] ... The mid-infrared (mid-IR) spectral region is the frequency range ...<|separator|>
  29. [29]
    Fourier Transform Spectrometry with Fourier Analysis of the ...
    Dec 26, 2018 · Fourier transform spectrometers replace the traditional dispersive frequency analyzer by a Michelson interferometer. The spectrum is the ...
  30. [30]
    Fourier Transform Infrared (FTIR) Spectroscopy as a Tool to ... - NIH
    Nov 20, 2024 · Thus, FTIR spectroscopy encounters two principal challenges that limit spectral accuracy. First, the presence of water molecules within samples ...
  31. [31]
    Atomic Spectroscopy - Chemistry and Biochemistry
    ... atom absorbs (a hollow cathode lamp), a flame or graphite furnace to heat the sample, a monochromator to select the wavelength of light, and a photodetector. In ...
  32. [32]
    [PDF] Chapter 9: Atomic Absorption Spectrometry - MSU chemistry
    Typical data for the determination of lead (Pb2+) by graphite furnace. AAS in ... hollow cathode lamp spectral interference atomization ionization ...
  33. [33]
    [PDF] Use of near infrared spectroscopy to measure the chemical and ...
    May 13, 2004 · These advantages include minimal sample preparation, rapid acquisition times, and non-contact, non- destructive spectral acquisition. Some ...
  34. [34]
    Examining the origins of observed terahertz modes from an optically ...
    The microscopic origins of terahertz (THz) vibrational modes in biological systems are an active and open area of current research.
  35. [35]
    [PDF] An Introduction to Processing, Fitting, and Interpreting Transient ...
    Feb 16, 2024 · Transient absorption (TA) spectroscopy is a time-resolved method tracking excited-state processes via changes in absorption spectrum after a ...
  36. [36]
    Detection Techniques for Lead Ions in Water: A Review - PMC
    Apr 20, 2023 · The detection limit of photometry is higher (0.01 mg/L), but this method can be achieved in most laboratories. The application of different ...
  37. [37]
    Monitoring Reactions Through UV-Visible Spectroscopy
    Aug 7, 2023 · UV-Visible absorption spectroscopy is a productive technique for monitoring reactions. The measured absorbance is directly proportional to the concentration of ...
  38. [38]
    Photometric complexometric titration - Metrohm
    May 25, 2020 · For a complexometric titration analysis, it is very important to know the qualitative composition of your sample. This determines the indicator, ...
  39. [39]
    (PDF) A Comprehensive Review on Hyphenated Techniques in ...
    Oct 9, 2025 · Hyphenated techniques in the field of analytical chemistry involve combining different analytical methods to enhance the analysis of samples.
  40. [40]
    The Classification of Stellar Spectra
    Joseph Fraunhofer made a more careful set of observations of the solar spectrum in 1814 and found some 600 dark lines, and he specifically measured the ...
  41. [41]
    [PDF] STELLAR SPECTRA A. Basic Line Formation
    Feb 6, 2007 · ... spectral lines in sunlight (“Fraunhofer lines”); their ... – Kirchhoff & Bunsen (1860): absorption lines in stellar spectra are the reverse of ...
  42. [42]
    [PDF] Chapter 3 The Interstellar Medium - UMD Astronomy
    There is H2 band absorption in the far ultraviolet, but this can only be observed from above the Earth's atmosphere.
  43. [43]
    [PDF] EXOPLANET TRANSIT SPECTROSCOPY USING WFC3: WASP-12 ...
    Our final transit spectra for all three objects are consistent with the presence of a broad absorption feature at 1.4 μm most likely due to water. However, the ...
  44. [44]
    DOAS Differential Optical Absorption Spectroscopy - IMPACT - NASA
    It detects trace gases by measuring specific, narrow-band absorptions in the ultraviolet and visible light spectrum. DOAS is typically used to detect ozone ...
  45. [45]
    Ozone Monitoring Instrument - NASA Earthdata
    Launched in 2004 aboard NASA's Aura satellite, OMI provides a record of total ozone and other atmospheric parameters related to ozone chemistry and climate ...
  46. [46]
    [PDF] Astronomy 404 September 27, 2013 Chapter 9. Stellar Atmospheres
    Sep 27, 2013 · Other Doppler broadenings involve stellar rotation, pulsation, and mass loss (P Cygni profile). 3. Pressure (and collisional) broadening. The ...Missing: spectroscopy | Show results with:spectroscopy
  47. [47]
    [PDF] Ground-based photon path measurements from solar absorption ...
    Retrieval of photon path lengths with sub-percentage precision from atmospheric spectra is always a challenge. For the A-band, the high optical depths mean ...
  48. [48]
    [PDF] Stellar Populations - Lecture III - UMD Astronomy
    ABUNDANCES FROM HIGH-RESOLUTION SPECTROSCOPY. With high-resolution spectra of individual stars, we can measure equivalent widths of absorption lines from ...
  49. [49]
    Helium first discovered during 1868 eclipse; the element later ...
    Mar 27, 2024 · The chemical element helium was first observed during an 1868 solar eclipse in India. French astronomer Pierre Janssen focused a spectroscope on the solar ...
  50. [50]
    Webb's Impact on Exoplanet Research - NASA Science
    Sep 2, 2025 · The James Webb Space Telescope has ushered in a new era in exoplanet research. To learn about the diversity of exoplanets and their ...
  51. [51]
    A Simple and General Method for Determining the Protein and ... - NIH
    Here, we consider the absorbance spectrum to be a sum of protein and nucleic acid components, determined algebraically from the UV absorption at 260nm and 280nm ...
  52. [52]
    Ultraviolet absorbance at 260 and 280 nm in RNA measurement is ...
    Aug 10, 2025 · The ratio of absorbance at 260 and 280 nm (the A260/280 ratio) is frequently used to assess the purity of RNA and DNA preparations. A low A260/ ...
  53. [53]
    Pulse oximetry - PMC - NIH
    Pulse oximetry is based on two physical principles: (a) the presenceof a pulsatile signal generated by arterial blood, which is relativelyindependent of non- ...Missing: spectroscopy | Show results with:spectroscopy
  54. [54]
    Point-of-care device to diagnose and monitor neonatal jaundice in ...
    As bilirubin concentration increased, measured spectra showed increasing absorbance in the blue with little to no change in absorption due to hemoglobin (Fig.
  55. [55]
    Advances in Clinical and Biomedical Applications of Photoacoustic ...
    Photoacoustic imaging is an imaging modality that derives image contrast from the optical absorption coefficient of the tissue being imaged.
  56. [56]
    UV-Vis Spectroscopy: Principle, Strengths and Limitations and ...
    Dec 18, 2023 · Samples contaminated with proteins will lower the 260/280 ratio due to higher absorbance at 280 nm. Wavelength used in absorbance analysis.<|control11|><|separator|>
  57. [57]
    Kinetic modelling: an integrated approach to analyze enzyme ... - NIH
    Aug 25, 2017 · For example, the spectrophotometric data of coupled reactions like the PDC assay initially showed a lag phase before reaching a linear phase ( ...
  58. [58]
    A new generation of sensors for non-invasive blood glucose ... - NIH
    This review compares the progress and challenges of electrochemical, optical, and electromagnetic/microwave non-invasive blood glucose monitoring techniquesElectrochemical Technology · Flexible Biosensors Based On... · Optical TechnologyMissing: portable 2020s
  59. [59]
    Optical spectroscopy for in vivo medical diagnosis—a review of the ...
    Aug 26, 2020 · In this review, several types of optical spectroscopy techniques (elastic optical scattering spectroscopy, fluorescence spectroscopy, Raman spectroscopy, and ...