Fact-checked by Grok 2 weeks ago

Spectroscopy

Spectroscopy is the scientific study of the between and , typically analyzed as a function of the or of the radiation emitted, absorbed, reflected, or scattered by the material. This field encompasses a wide range of techniques used to identify substances, determine their composition, and probe their physical and chemical properties by examining the unique signatures produced by and molecular transitions. The history of spectroscopy traces back to the 17th century, when Sir Isaac Newton demonstrated that white light could be dispersed into a continuous of colors using a , laying the foundational understanding of light's composition. Significant advancements occurred in the , with Fraunhofer's identification of dark absorption lines in the solar in 1814, and and Robert Bunsen's formulation in 1859 of the principle that every element produces a unique , enabling the first spectrochemical analyses of celestial bodies. The discovery of radiation by in 1800 and ultraviolet by Johann Ritter in 1801 expanded the spectral range beyond visible light. The marked the era of modern spectroscopy, propelled by and the invention of the in 1960, which provided intense, monochromatic sources for precise measurements. At its core, spectroscopy operates on the principle that atoms and molecules absorb or emit at specific wavelengths corresponding to quantized transitions, as described by . Common techniques include emission spectroscopy, in which excited atoms release to produce bright lines against a dark background, and , which reveals dark lines in a continuous where light is absorbed. Types of spectra include continuous spectra, produced by hot, dense objects like emitting across all wavelengths. Reflectance and processes further contribute to , particularly in solid or particulate materials, where and composition influence the observed signatures. Spectroscopy finds broad applications across disciplines, from determining the , temperature, and velocity of astronomical objects via and Doppler shifts in , to analyzing molecular structures in and through techniques like and . In materials science and , it identifies minerals and environmental contaminants by their reflectance properties. These methods are indispensable for non-destructive analysis, enabling insights into everything from exoplanet atmospheres to pharmaceutical .

Fundamentals

Definition and Scope

Spectroscopy is the branch of science that studies the interaction of matter with through processes such as , , or , which produce a that provides insights into the and molecular , , and of the under . This field enables the qualitative and quantitative analysis of substances by examining how they respond to different wavelengths of , revealing unique spectral signatures that characterize their levels and chemical bonds. The term "spectroscopy" originates from the Latin word , meaning "image" or "apparition," combined with the Greek skopia, meaning "observation" or "examination," reflecting its focus on observing spectral phenomena. Although the systematic study of light dispersion began in the 17th century with Isaac Newton's work on prisms, the modern term was coined in the mid-19th century amid advances in by scientists like and . The scope of spectroscopy encompasses the entire , from low-energy radio waves with wavelengths on the order of meters to high-energy gamma rays with wavelengths shorter than atomic nuclei, allowing applications across diverse scales from subatomic particles to astronomical objects. It is important to distinguish spectroscopy from related terms: while spectroscopy broadly refers to the study and interpretation of spectra, spectrometry emphasizes the measurement of these spectra using instruments, and specifically involves the quantitative measurement of light absorption or transmission as a function of for analytical purposes. Central to spectroscopic analysis are different types of spectra: a continuous spectrum appears as a smooth distribution of wavelengths, like that from a hot incandescent source; an features bright lines or bands against a dark background, indicating wavelengths emitted during electronic transitions; and an absorption spectrum shows dark lines superimposed on a continuous spectrum, corresponding to wavelengths absorbed by the sample. Two fundamental metrics in spectroscopy are and (SNR). , often quantified as R = \frac{\lambda}{\Delta \lambda}, where \lambda is the and \Delta \lambda is the smallest resolvable difference, determines the instrument's ability to distinguish closely spaced spectral features, such as fine atomic lines. SNR, defined as the ratio of the signal intensity to the noise level (typically the root-mean-square of background fluctuations), assesses the quality and reliability of spectral data, with higher values enabling clearer detection of weak signals amid noise. These metrics are crucial for ensuring the precision and interpretability of spectra across the field's interdisciplinary applications in , physics, , and astronomy.

Basic Principles of Radiation-Matter Interaction

, the basis of all spectroscopic techniques, is characterized by its \lambda, \nu, and energy E. The and are inversely related through the c in , given by \nu = c / \lambda. The energy of a , the quantum of electromagnetic radiation, is directly proportional to its via Planck's relation E = h\nu, where h is Planck's constant ($6.626 \times 10^{-34} J s), highlighting the quantized nature of . This relation underscores the wave-particle duality of radiation, which behaves as both waves (exhibiting and ) and particles (photons with discrete energy packets) depending on the experimental context. The primary mechanisms of interaction between and matter in spectroscopy are , , and . occurs when a is taken up by an or , promoting an to a higher or exciting a vibrational mode, provided the 's matches the difference between . is the reverse process, where an excited species relaxes to a lower , releasing a of corresponding ; this is observed in techniques like or . involves the redirection of without net , divided into elastic scattering (e.g., , where the 's remains unchanged) and inelastic scattering (e.g., , where the scattered gains or loses due to vibrational or rotational transitions in the ). For absorption processes, the extent of through a medium is quantified by the Beer-Lambert law, which describes the of intensity. The law derives from considering the incremental loss of photon flux dI over a path length dz due to by N absorbers each with cross-section \sigma, yielding dI / dz = -\sigma N I, or integrated form I = I_0 e^{-\alpha l} where \alpha = \sigma N is the coefficient and l is the path length. In terms of concentration c (moles per ), this becomes the A = -\log_{10}(I / I_0) = \epsilon c l, where \epsilon is the molar absorptivity (a measure of the absorber's efficiency at a given ). This linear relationship enables of concentrations in spectroscopic measurements. Not all transitions between energy states are equally probable; selection rules dictate which interactions are allowed based on quantum mechanical constraints. These rules arise from the properties of the wavefunctions and the interaction operator in time-dependent , requiring the integral \langle \psi_2 | \hat{\mu} | \psi_1 \rangle to be non-zero for observable transitions. For electric transitions, common in UV-Vis and spectroscopy, allowed changes in quantum numbers include \Delta l = \pm 1 for orbital and specific or matches (e.g., gerade to ungerade for homonuclear diatomics). Vibrational selection rules, such as \Delta v = \pm 1 for harmonic oscillators, further restrict observations to modes that alter the molecular . These rules ensure that only certain spectral lines appear, providing insights into molecular structure and .

Theoretical Foundations

Quantum Mechanical Basis

The quantum mechanical foundation of spectroscopy rests on the time-independent , which describes the stationary states of bound . For a particle in a potential V(\mathbf{r}), the equation takes the form \hat{H} \psi = [E](/page/E!) \psi, where \hat{H} = -\frac{\hbar^2}{2m} \nabla^2 + V(\mathbf{r}) is the operator, \psi is , and [E](/page/E!) is the eigenvalue. This yields discrete levels for bound systems, such as electrons in atoms or nuclei, enabling the prediction of spectral transitions between these levels when the system interacts with . In atomic systems, solving the Schrödinger equation for hydrogen-like atoms produces quantized energy levels given by E_n = -\frac{13.6 \, \mathrm{eV}}{n^2}, where n is the principal quantum number; this formula arises from the Coulomb potential and spherical symmetry, with higher-Z atoms scaled by Z^2. For molecules, the Born-Oppenheimer approximation separates the fast electronic motion from slower nuclear vibrations and rotations, leading to discrete vibrational energy levels approximated as E_v = \hbar \omega (v + \frac{1}{2}) for harmonic oscillators (with quantum number v and frequency \omega), combined with rotational levels E_J = \frac{\hbar^2}{2I} J(J+1) (where J is the rotational quantum number and I the moment of inertia). These quantized levels underpin the discrete spectral features observed in absorption and emission spectra. Spectroscopic transitions occur via time-dependent perturbations from the , treated using first-order . In the electric , valid for wavelengths much larger than the system size, the interaction is \hat{H}' = -\mathbf{\mu} \cdot \mathbf{E}, where \mathbf{\mu} is the operator and \mathbf{E} the . The transition rate from initial state |i\rangle to final state |f\rangle is given by : \Gamma = \frac{2\pi}{\hbar} |\langle f | \hat{H}' | i \rangle|^2 \rho(E_f - E_i), with \rho(E) the density of final states; this yields the probability per unit time for or when is conserved. The strength of such dipole-allowed transitions is quantified by the f_{ij} = \frac{8\pi^2 m \nu}{3 h e^2} |\mu_{ij}|^2, where \nu is the transition frequency, m the , e the charge, and \mu_{ij} = \langle i | \mathbf{\mu} | j \rangle; this dimensionless measure relates quantum probabilities to classical oscillator models and determines line intensities in spectra.

Spectral Lines and Transitions

Spectral lines arise from quantum mechanical transitions between discrete levels in atoms or molecules, where the or of photons at specific frequencies corresponds to the difference ΔE = hν between the levels, as predicted by . These lines are not infinitely sharp but exhibit finite widths and shapes due to various physical processes that perturb the ideal delta-function response. The observed line profile reflects the of multiple broadening mechanisms, influencing the resolution and interpretation of spectra in experimental settings. The primary intrinsic broadening is natural broadening, stemming from the finite lifetime τ of the due to . According to the time-energy , this leads to an energy uncertainty ΔE ≈ ħ / τ, manifesting as a line shape in the with (FWHM) Γ = 1/τ in angular frequency units, or Δν = 1/(2πτ) in frequency units. This profile is symmetric and has extended wings, described by the L(ν) = (Γ / 2π) / [(ν - ν₀)² + (Γ / 2)²], where ν₀ is the central ; it dominates in low-pressure environments where other effects are minimal. Thermal motion of atoms or molecules in a gas introduces , arising from the relativistic frequency shift δν / ν = v / c along the , where v is the component. The Maxwell-Boltzmann distribution yields a Gaussian line shape with FWHM Δν = (ν / c) √( (8 k T ln 2) / M ), where k is Boltzmann's constant, T is , and M is the atomic or . This broadening increases with and decreases with mass, often convolving with the to form a in real spectra. The intensity of spectral lines, quantified by the integrated absorption coefficient ∫ α(ν) dν, is proportional to the transition dipole moment squared |μ_{if}|² and the population difference between initial and final states, as given by the Einstein coefficients for absorption. For molecular spectra involving vibrational changes, the Franck-Condon principle governs the overlap of vibrational wavefunctions between electronic states, determining relative intensities through factors |⟨χ_v' | χ_v'' ⟩|², where χ denotes vibrational wavefunctions; vertical transitions with maximal overlap yield the strongest lines, explaining progressions in band spectra. External fields further modify line positions and structures. The causes splitting in a B due to the interaction of the with the field, with energy shift ΔE = μ_B g m_j B, where μ_B is the , g is the , and m_j is the ; this results in linearly or circularly polarized components, enabling diagnostics. Similarly, the in an E induces splitting via the quadratic Stark shift ΔE ∝ α E² (for non-degenerate states, with α) or linear shifts in degenerate cases like , altering line positions and widths proportionally to E. Hyperfine structure emerges from the magnetic dipole and electric quadrupole interactions between the nuclear spin I and the total electron angular momentum J, forming total angular momentum F = I + J with energy shifts A [F(F+1) - I(I+1) - J(J+1)] / 2, where A is the hyperfine coupling constant. This fine splitting, typically on the order of MHz to GHz, arises primarily from the Fermi contact term for s-electrons and dipolar interactions, resolving nuclear properties in high-resolution spectra.

Classification of Methods

By Radiative Energy Type

Spectroscopic methods are classified by the type of radiative energy employed, which corresponds to distinct regions of the . Each region targets specific energy differences in matter, from low-energy spin and rotational transitions to high-energy core and nuclear excitations. This classification highlights how the matches the scale of intramolecular or processes, enabling selective probing of , electronic structures, and nuclear properties. In the radio and microwave regions (frequencies typically 10 MHz to 300 GHz, corresponding to energies of approximately 4 \times 10^{-8} to 10^{-3} ), spectroscopy focuses on low-energy transitions involving and spins as well as molecular rotations. (NMR) spectroscopy uses radio-frequency pulses to induce transitions between nuclear spin states in a , providing detailed information on molecular connectivity and environments. spin resonance (ESR), also known as (), employs radiation to excite spin transitions, particularly useful for studying paramagnetic and free radicals. in the microwave domain measures pure rotational transitions, yielding molecular moments of inertia that reveal bond lengths and angles in gas-phase molecules. Infrared (IR) spectroscopy utilizes mid- to far-IR (wavelengths 2.5–50 μm, energies ~0.025–0.5 eV), which matches the energy scale of molecular vibrational modes. or in this region excites , , and other , allowing identification of functional groups through characteristic band positions. The region, spanning 400–1500 cm⁻¹, is particularly diagnostic as it encodes unique patterns of molecular , enabling compound identification akin to a "molecular ." Visible and ultraviolet (UV-Vis) spectroscopy covers wavelengths from 10–780 nm (energies ~1.6–124 eV), aligning with electronic transitions in valence shells. These methods probe promotions of electrons between molecular orbitals, such as π–π* or n–π* transitions in organic compounds. In transition metal complexes, d–d bands arise from splitting of d-orbitals in ligand fields, providing insights into coordination geometry and oxidation states. X-ray and gamma-ray spectroscopy employs high-energy photons (energies >100 eV for X-rays and >100 keV for gamma rays), targeting inner-shell processes. X-ray absorption spectroscopy excites core electrons to higher levels, revealing local electronic structure and oxidation states around atoms. Mössbauer spectroscopy, using gamma rays, examines recoilless nuclear transitions in solids, sensitive to the nuclear environment and isomer shifts indicative of chemical bonding. These radiative energy regions correspond to hierarchical energy scales in : rotational and transitions at micro- to milli-electronvolt (μeV–meV) levels (radio/), vibrational modes at milli- to fraction-of-electronvolt (meV–~0.5 ) (IR), valence electronic transitions at () scales (UV-Vis), core-level excitations at kiloelectronvolt (keV) (X-rays), and nuclear levels at higher keV (gamma rays). This matching ensures specificity in probing different physical phenomena.

By Interaction Nature

Spectroscopic methods can be classified by the primary interaction between and , encompassing , , , and coherence-based processes. This emphasizes the fundamental physical processes involved, such as or alterations, rather than the specific energy range of the . These interactions enable the probing of molecular, , or material properties through distinct signatures in the detected signals. Absorption spectroscopy measures the intensity of light transmitted through a sample after interaction with matter, where the reduction in transmitted light intensity quantifies the at specific wavelengths. This technique directly probes transitions from to excited states, with the absorbed intensity proportional to the population in the lower energy state, governed by the Beer-Lambert law: A = \epsilon c l, where A is , \epsilon is the , c is concentration, and l is path length. Emission spectroscopy involves exciting a sample to higher states and analyzing the emitted as it relaxes back to lower states, providing information on excited-state dynamics. The emitted reveals levels and transition probabilities, often through or . A key parameter is the fluorescence lifetime \tau, defined as the average time a spends in the before emission, given by \tau = \frac{1}{k_r + k_{nr}}, where k_r is the radiative decay rate and k_{nr} is the non-radiative decay rate; this lifetime is sensitive to environmental factors like . Scattering spectroscopy distinguishes elastic and inelastic processes based on energy exchange with matter. Rayleigh scattering is elastic, where incident photons are redirected without frequency change (\Delta \nu = 0), arising from temporary of the sample. In contrast, is inelastic, with scattered light shifted by \Delta \nu = \nu_{vib}, corresponding to molecular vibrational frequencies, enabling vibrational spectroscopy without . Brillouin scattering, another inelastic variant, involves interactions with acoustic phonons, producing frequency shifts on the order of GHz that probe sound velocities and elastic properties in materials. Coherence methods, such as , exploit relationships in the radiation-matter interaction for enhanced detection. By superimposing reference and sample beams in a Michelson or Mach-Zehnder interferometer, shifts induced by the sample's or dispersion are measured, enabling high-resolution, phase-sensitive spectroscopy. This approach improves signal-to-noise ratios for weak interactions, as seen in where interferograms are transformed to reveal spectral details. Detection limits in these methods are fundamentally tied to the interaction cross-sections, which quantify the probability of photon-matter events per molecule. Larger cross-sections, as in vibrational transitions, enhance for trace analytes, achieving limits down to parts-per-billion in optimized setups; for instance, IR absorption can detect sub-μg/mL concentrations via pathlength scaling. Noise sources like detector limit ultimate , but cross-section-driven signal strength allows selective tracing of analytes in complex matrices.

Spectroscopic Techniques by Target

Atomic Spectroscopy

Atomic spectroscopy examines the spectra produced by free atoms and ions in the gas phase, where interactions with electromagnetic radiation reveal the electronic structure through discrete transitions between quantized energy levels. These transitions generate sharp, well-defined spectral lines, each corresponding to a specific energy difference unique to the , enabling precise identification and quantification without the broadening effects from molecular vibrations or rotations. The simplicity of atomic spectra arises from the isolated nature of atoms, making this branch essential for analyzing gaseous samples and vaporized materials. In , energy levels are classified using spectroscopic s under the LS (Russell-Saunders) coupling approximation, suitable for lighter atoms where orbital and angular momenta couple separately before combining. The is denoted as ^{2S+1}L_J, where L represents the total orbital angular momentum (with letters S for L=0, P for L=1, D for L=2, etc.), S is the total , $2S+1 is the multiplicity indicating spin degeneracy, and J is the total angular momentum ranging from |L - S| to L + S. Transitions between these levels obey selection rules, such as \Delta L = \pm 1 for electric dipole radiation, producing the observed sharp lines in or spectra. This notation facilitates the prediction of spectral patterns and intensities for multi-electron atoms. A foundational technique in atomic spectroscopy is atomic absorption spectroscopy (AAS), which quantifies elements by measuring the absorption of resonant light by ground-state atoms in a vaporized sample. The process involves atomization—typically via flame or graphite furnace—to produce free atoms, followed by passage through a hollow cathode lamp emitting element-specific wavelengths; the degree of absorption follows Beer's law, where intensity is proportional to atomic concentration. AAS excels in trace elemental analysis, detecting metals at parts-per-billion levels in environmental, clinical, and industrial samples, though interferences from molecular species or matrix effects require corrections like background subtraction. Flame photometry, a form of , is optimized for and alkaline metals, exploiting their low excitation energies to produce intense emission lines when aspirated into a . In this method, the sample solution is nebulized into a propane-air or acetylene-air , where evaporates, and atoms are thermally excited; emitted is isolated via filters and measured photometrically, with curves linking intensity to concentration. It is particularly sensitive for sodium (589 nm doublet) and (766 nm), achieving detection limits around 0.1 ppm, and finds routine use in , , and biological fluid analysis for these ions. For simultaneous multi-element detection, optical emission spectroscopy (ICP-OES) employs a high-temperature (around 6000–10,000 K) generated by radio-frequency to fully atomize and ionize samples, exciting atoms to emit lines across the UV-visible range. The 's robustness allows analysis of matrices like geological waters or digests, detecting over 70 at trace levels (e.g., 1–10 ppb for many metals) using axial or radial viewing configurations for optimal signal-to-noise. interferences, where overlapping emissions from different obscure signals (e.g., iron lines near ), are resolved through high-resolution echelle gratings, correction algorithms, and inter-element correction factors derived from additions. Official methods like EPA 6010C standardize ICP-OES for , emphasizing matrix matching and quality controls to ensure accuracy. In astrophysics, atomic spectroscopy underpins elemental abundance determinations in stellar atmospheres, exemplified by the Fraunhofer absorption lines in the Sun's visible spectrum, which arise from cooler gas layers absorbing continuum radiation at atomic transition wavelengths. These dark lines, cataloged extensively since the 19th century, match laboratory spectra of elements like hydrogen (Balmer series at 656 nm, 486 nm), calcium (K line at 393 nm), and iron, revealing the solar photosphere's composition (e.g., ~74% hydrogen by mass). Advances in solar spectroscopy, including ultraviolet observations, have refined these identifications, confirming atomic processes dominate the line formation in low-density plasmas.

Molecular Spectroscopy

Molecular spectroscopy probes the energy levels associated with vibrational, rotational, and motions in molecules, providing insights into their structure, bonding, and dynamics. These introduce complexity beyond systems, as molecular spectra often exhibit from coupled transitions. Vibrational and rotational spectra appear in the and regions, while electronic spectra dominate the ultraviolet-visible range, with overlaps enabling detailed characterization of polyatomic and diatomic in isolation or dilute phases. Rovibrational spectra, observed primarily in the , arise from simultaneous changes in vibrational and rotational quantum numbers, resulting in banded structures. For diatomic molecules, the fundamental vibrational transition (Δv = +1) is accompanied by rotational changes governed by selection rules ΔJ = ±1 or 0, producing distinct P, , and R branches. The P branch (ΔJ = -1) appears at lower frequencies than the band origin, corresponding to transitions from higher initial rotational levels to lower final levels; the R branch (ΔJ = +1) shifts to higher frequencies; and the branch (ΔJ = 0), when allowed, forms a , though it is often absent in simple diatomics due to restrictions. These branches enable of rotational constants B and vibrational frequencies from spacing and band contours. Real molecular vibrations deviate from the model due to in the , leading to non-equispaced energy levels and observable . The anharmonic oscillator energy levels are approximated by E_v = h\nu \left( v + \frac{1}{2} \right) - h\nu x_e \left( v + \frac{1}{2} \right)^2 where v is the vibrational , \nu is the harmonic frequency, and x_e > 0 is the anharmonicity constant (small, typically 0.01-0.03). This negative correction accounts for weaker higher (Δv > 1) and interactions between levels, improving spectral predictions for polyatomics. Electronic spectroscopy in the UV-Vis region excites valence electrons between molecular orbitals, often revealing conjugation and functional groups. The π → π* transition, prominent in unsaturated hydrocarbons and aromatics, involves promotion from a bonding π orbital to an antibonding π* orbital, absorbing around 200–400 nm depending on conjugation length; for example, shows a strong band near 175 nm./New_Page/4%3A_Structure_Determination_I-_UV-Vis_and_Infrared_Spectroscopy_Mass_Spectrometry/4.4%3A_Ultraviolet_and_visible_spectroscopy) The depicts these processes, mapping singlet (S) and triplet (T) states with vertical electronic transitions and horizontal non-radiative relaxations. Singlet-triplet (ISC) from S₁ to T₁ occurs via spin-orbit coupling, enabling or energy transfer, with rates enhanced in heavy-atom substituted molecules. Fourier transform infrared (FTIR) spectroscopy is a cornerstone technique for rovibrational analysis, offering high sensitivity and resolution for gas-phase studies like atmospheric monitoring and liquid-phase identification of solvents or biomolecules. In gases, long path lengths exploit weak absorptions for trace detection; in liquids, (ATR) modes minimize preparation. , based on changes, complements FTIR by providing orthogonal selection rules and is particularly suited to aqueous environments, where water's weak scattering avoids the strong IR absorption bands that interfere with studies. Isotopic substitution alters spectral features through changes in the μ = m₁m₂/(m₁ + m₂), scaling vibrational frequencies as ν ∝ √(k/μ), where k is the force constant. Heavier isotopes increase μ, lowering band positions; for instance, replacing ¹H with ²H in HCl shifts the fundamental from ~2990 cm⁻¹ to ~2145 cm⁻¹, aiding mode assignment and confirming molecular identity in complex mixtures. These shifts are more pronounced for vibrations involving the substituted atom, enabling for dynamic studies.

Condensed Matter Spectroscopy

Condensed matter spectroscopy investigates the electronic, vibrational, and structural properties of solids and extended materials, emphasizing collective excitations such as phonons and plasmons, as well as band structures that arise from periodic lattices. Unlike isolated systems, these materials exhibit delocalized states and interactions that lead to emergent phenomena observable through spectroscopic probes. This field is crucial for understanding semiconductors, metals, and insulators, where spectroscopy reveals how orbitals form energy bands influencing and optical responses. In band theory, the periodic potential of a crystal lattice causes wavefunctions to form Bloch states, resulting in bands separated by s. Semiconductors feature a small , typically 0.1 to 4 eV, where the valence band (filled with s) and conduction band (empty) play key roles in charge transport. Direct s occur when the valence band maximum and conduction band minimum align at the same wavevector k in the , enabling efficient radiative recombination, as in GaAs with a gap of 1.42 eV at . In contrast, indirect s, like in (1.12 eV), require assistance for momentum conservation during transitions, reducing luminescence efficiency. Excitons in semiconductors are bound electron-hole pairs formed when an is excited from the to the conduction band, held together by attraction with binding energies of 10-100 meV in typical materials. These quasi-particles behave like atoms but are influenced by the screening of the host lattice, leading to Wannier-Mott excitons in three-dimensional semiconductors with large radii (up to hundreds of angstroms). In direct-gap materials, excitons manifest as sharp lines below the band edge, enhancing in devices like LEDs. X-ray photoelectron spectroscopy (XPS) probes surface composition in condensed matter by measuring the kinetic energy of photoelectrons emitted from core levels under irradiation, with penetration depths of 1-10 . Binding energy shifts, arising from chemical environment changes (e.g., oxidation states), allow identification of elemental ; for instance, carbon 1s peaks shift by 1-5 eV depending on bonding. This technique is essential for analyzing surface oxides or interfaces in materials like thin films. Ultraviolet photoelectron spectroscopy (UPS) examines valence band structures in solids using helium discharge lamps (photon energies 21.2-40.8 eV), ejecting electrons from occupied states near the . It maps the , revealing band dispersions and work functions; for example, in metals, it shows a broad valence band width of 5-10 eV. UPS complements XPS by focusing on delocalized electrons critical for electronic properties in semiconductors and organics. Phonon spectroscopy in solids uses (IR) absorption and to probe vibrations, which are collective modes representing quantized displacements of atoms. IR spectroscopy detects polar phonons that couple to electromagnetic fields, exciting transverse optical (TO) modes at frequencies around 10-15 THz in ionic crystals like NaCl. , involving inelastic light scattering, reveals both optical and acoustic phonons through changes in , often showing longitudinal optical (LO) modes split from TO by the Lyddane-Sachs-Teller relation. These techniques identify symmetries and in materials. Phonon dispersion relations describe how vibrational frequencies depend on wavevector, given by \omega(\mathbf{k}), where acoustic branches start at zero frequency at the Brillouin zone center (k=0) and flatten near zone boundaries due to short-range forces. In one-dimensional chains, the relation is \omega(k) = 2\sqrt{\frac{K}{m}} \left| \sin\left(\frac{ka}{2}\right) \right| for nearest-neighbor springs of constant K and m, illustrating the transition from sound-wave-like to standing waves. In three-dimensional crystals, neutron scattering maps these relations, revealing optic-acoustic gaps essential for thermal conductivity. Defects and impurities in condensed matter introduce localized states within the band gap, leading to mid-gap features in optical spectra. These states, such as donor levels 0.01-0.1 below the conduction band in doped , trap carriers and cause sub-bandgap transitions observable via or . In wide-gap materials like , deep impurities form levels at mid-gap (e.g., 2-3 ), acting as recombination centers that broaden tails and reduce efficiency in . Spectroscopy distinguishes these via characteristic peak positions and linewidths.

Nuclear Spectroscopy

Nuclear spectroscopy encompasses techniques that probe the intrinsic properties of atomic nuclei, such as their , , and structures, primarily through interactions with external s or gamma radiation. These methods reveal information about levels, which are influenced by factors like the nuclear —a measure of the deviation from spherical symmetry in the nuclear charge distribution for nuclei with I > 1/2. For instance, nuclear quadrupole resonance (NQR) spectroscopy detects transitions between these levels in the absence of a magnetic field, providing insights into the at the . In , recoilless emission and absorption of gamma rays from nuclei like ⁵⁷ allow precise measurement of shifts, which reflect changes in the nuclear s-electron density due to the chemical environment; typical shifts for low- () compounds range from -0.1 to 0.2 mm/s relative to α-iron. A cornerstone of spectroscopy is (NMR), where nuclei with nonzero precess in a at the Larmor frequency, given by \omega = \gamma B, with \gamma as the specific to each and B as the applied magnetic field strength—for protons, \gamma \approx 2.675 \times 10^8 rad/s/T. This precession leads to resonant absorption of radiofrequency energy, the of states. The , quantifying the in to shielding, is defined as \delta = \frac{\nu_\text{sample} - \nu_\text{ref}}{\nu_\text{ref}} in parts per million (ppm), allowing differentiation of environments in molecules. Key techniques in nuclear spectroscopy include solid-state NMR, which overcomes the broadening effects of anisotropic interactions in rigid samples like polymers to characterize chain dynamics and morphology; for example, it reveals and crystallinity in through ¹³C spectral lineshapes. (EPR) spectroscopy complements this by detecting paramagnetic centers, such as ions or organic radicals, where unpaired electrons interact with nearby nuclei. Hyperfine interactions in these systems are described by the Hamiltonian term H = A \mathbf{I} \cdot \mathbf{S}, with A as the between nuclear spin \mathbf{I} and electron spin \mathbf{S}, manifesting as splitting patterns that provide structural details around the paramagnetic site.

Advanced and Specialized Methods

Time-Resolved Spectroscopy

Time-resolved spectroscopy encompasses techniques designed to observe and characterize dynamic processes in materials and molecules on timescales ranging from femtoseconds to seconds, providing insights into transient states that are inaccessible through steady-state measurements. These methods typically involve perturbing a sample with an excitation pulse and monitoring its evolution using subsequent probe pulses or detection schemes, capturing phenomena such as , charge separation, and structural rearrangements. By resolving temporal evolution, time-resolved spectroscopy reveals the of excited-state processes, enabling the study of ultrafast events like vibrational relaxation and transitions. A cornerstone technique is , which employs ultrashort laser on the femtosecond to picosecond scale to initiate and interrogate dynamics. In this approach, a "" excites the sample to a non-equilibrium state, such as populating an electronic , while a delayed "" measures the resulting changes in , transmission, or reflection. The transient signal, denoted as \Delta A(t), quantifies the difference in probe intensity with and without the pump, where t is the time delay between ; positive \Delta A indicates increased due to excited-state species, while negative values signify ground-state bleaching or . This method achieves sub-picosecond , limited primarily by the duration and instrument response function, and is widely used to track processes like solvation dynamics and cooling. Femtochemistry, pioneered through pump-probe experiments, applies these ultrafast timescales to dissect pathways at the atomic level, particularly bond breaking and formation. In this field, reactions are "filmed" by resolving transient intermediates, such as those traversing s—funnel-like regions on surfaces where electronic states couple, enabling rapid nonradiative decay from excited to ground states on timescales. For instance, bond dissociation in molecules like ICN occurs via a conical intersection with lifetimes \tau \approx 200 fs, allowing direct observation of the state's and dissipation. These studies, conducted with , have elucidated mechanisms in and predissociation, demonstrating how vibrational coherence drives selectivity in reaction outcomes. Another key method is time-correlated single photon counting (TCSPC), which excels in measuring fluorescence decay kinetics for longer-lived excited states on picosecond to nanosecond scales. TCSPC operates by exciting the sample with repetitive short pulses and recording the arrival times of individual emission photons relative to each excitation event, building a histogram of time delays to reconstruct the decay profile. The technique relies on single-photon detectors with low jitter (e.g., <50 ps) and ensures statistical accuracy by maintaining a low photon detection probability per cycle (<0.05), enabling deconvolution of multi-exponential decays associated with heterogeneous environments or energy transfer. TCSPC provides high signal-to-noise ratios for weak signals, making it ideal for probing radiative lifetimes in biomolecules and nanomaterials. Applications of time-resolved spectroscopy span photochemical reactions, where it tracks ultrafast electron and proton transfer in systems like photoexcited dyes, revealing pathways for generation with reaction times under 100 . In solar cells, particularly perovskites, these techniques monitor carrier , including hot carrier cooling (on ~1 ps scales) and recombination at interfaces, which dictate device ; for example, transient has quantified charge extraction times of 10-100 ps in high-performance cells, guiding material optimizations. Such insights have advanced understanding of photovoltaic losses and photochemical , bridging with practical device performance.

Nonlinear Spectroscopy

Nonlinear spectroscopy encompasses techniques that exploit high-intensity light fields to elicit material responses proportional to higher powers of the , enabling the study of subtle interactions and dynamics beyond linear regimes. These methods rely on nonlinear optical susceptibilities, such as the second-order χ^{(2)} for processes involving even-order field terms and third-order χ^{(3)} for odd-order ones, which arise in non-centrosymmetric media or through multi-photon effects. By using intense pulses, nonlinear spectroscopy probes vibrational, electronic, and orientational properties with enhanced selectivity and sensitivity. Second-harmonic generation (SHG) exemplifies a χ^{(2)} , where two incident of frequency ω interact to produce a scattered photon at 2ω, described by the : \mathbf{P}^{(2)}(\omega=2\omega_0) = \epsilon_0 \chi^{(2)} : \mathbf{E}(\omega_0) \mathbf{E}(\omega_0) This coherent, directional emission was first demonstrated in 1961 using a focused into crystals. SHG is particularly useful for characterizing non-centrosymmetric structures, such as chiral molecules or surfaces, as the process vanishes in centrosymmetric environments due to constraints. Two-photon absorption (TPA) represents another key χ^{(2)}-like effect, where the excitation rate to a higher electronic state scales quadratically with intensity: Γ = σ_2 I, with σ_2 the two-photon cross-section (typically in Göppert-Mayer units, GM = 10^{-50} cm^4 s photon^{-1}) and I the beam intensity. Observed experimentally in 1961 in Eu^{2+}-doped CaF_2 crystals using pulses, TPA allows access to excited states that are symmetry-forbidden for single-photon transitions, such as certain S_0 to S_1 promotions in organic dyes. This enables deeper penetration in near-infrared wavelengths compared to linear . Advanced techniques like two-dimensional (2D-) spectroscopy extend these principles to vibrational domains, employing sequences of pulses to generate χ^{(3)} signals that map frequency correlations. In 2D-, the amide I band (around 1600–1700 cm^{-1}, arising from C=O stretches in peptides) reveals secondary motifs in proteins through cross-peak intensities indicating between delocalized . Comprehensive reviews highlight its application in resolving congestion and dynamics in biomolecules. Coherent anti-Stokes Raman scattering (CARS) is a prominent χ^{(3)}-based method for label-free vibrational spectroscopy, involving pump (ω_p), Stokes (ω_s), and probe (often ω_p) beams to generate an anti-Stokes signal at ω_as = 2ω_p - ω_s, resonant with molecular vibrations. First reported in 1965 through studies of third-order polarization in liquids, provides rapid, chemically specific imaging without fluorescent labels, leveraging the Raman resonance for high signal-to-noise. Four-wave mixing (FWM) generalizes these interactions, where three input waves at frequencies ω_1, ω_2, ω_3 produce a fourth at ω_4 = ω_1 + ω_2 - ω_3, governed by the χ^{(3)} tensor. Efficient signal generation requires phase-matching to minimize wavevector mismatch, satisfying conditions like \mathbf{k_1} + \mathbf{k_2} = \mathbf{k_s} + \mathbf{k_i} for signal (s) and idler (i) outputs, ensuring constructive over the interaction length. Reviews emphasize FWM's role in ultrafast spectroscopy for tracking coherences and . These nonlinear approaches offer distinct advantages, including background-free detection from coherent, phase-coherent emission that rejects non-resonant fluorescence, and the ability to probe forbidden transitions via multi-photon pathways. For instance, CARS signals are inherently directional and free of solvent interference, enhancing contrast in complex samples.

Computational Spectroscopy

Computational spectroscopy encompasses the use of theoretical and numerical methods to simulate and predict molecular spectra, enabling the interpretation of experimental data and the exploration of systems inaccessible to direct measurement. These approaches rely on quantum mechanical frameworks to compute electronic, vibrational, and rotational energy levels, from which spectral features such as absorption bands and intensities are derived. By bridging the gap between atomic-scale quantum mechanics and observable spectroscopic signatures, computational methods facilitate the design of novel materials and the elucidation of reaction mechanisms. A cornerstone of computational spectroscopy for electronic spectra is time-dependent density functional theory (TD-DFT), which efficiently calculates excitation energies and oscillator strengths for medium-sized molecules. TD-DFT extends ground-state DFT to time-dependent perturbations, providing vertical transition energies that approximate UV-Vis and fluorescence spectra with reasonable accuracy for many organic and inorganic systems. For higher precision in energy calculations, particularly for benchmark studies of small molecules, the coupled-cluster method with single, double, and perturbative triple excitations, CCSD(T), serves as a gold standard, yielding near-quantitative agreement with experimental ionization potentials and electron affinities relevant to photoelectron spectroscopy. These ab initio techniques often employ correlation-consistent basis sets to approach the complete basis set limit, ensuring reliable predictions of spectral shifts due to electronic correlations. Spectral simulations further refine these predictions by incorporating vibronic and anharmonic effects. Franck-Condon factors, which govern the intensity distribution in vibrational progressions of electronic spectra, are computed using vertical gradients of excited-state surfaces, allowing the reconstruction of envelopes without full multidimensional potential scans. For vibrational spectroscopy, corrections address deviations from the model, using methods like vibrational self-consistent field (VSCF) theory to shift fundamental frequencies and adjust overtone intensities, thus improving matches to observed and Raman bands in polyatomic molecules. These simulations often integrate over Boltzmann-distributed conformations to yield temperature-dependent spectra. Recent advances incorporate , particularly neural networks, to accelerate prediction directly from molecular structures, bypassing costly quantum calculations for large datasets. Graph neural networks or message-passing architectures trained on quantum-derived spectra can predict or NMR features with errors below 5 cm⁻¹ for vibrational modes, enabling in and materials design. Such models learn implicit mappings from to spectral fingerprints, enhancing efficiency while maintaining chemical interpretability through attention mechanisms. Despite these strengths, computational spectroscopy faces limitations from approximations in basis sets and environmental modeling. Basis set superposition error (BSSE) artificially stabilizes intermolecular interactions in weakly bound complexes, inflating binding energies in simulated Raman spectra of clusters; counterpoise corrections mitigate this by evaluating fragments with ghost orbitals. , crucial for solution-phase spectra, are often treated with polarizable continuum models (PCM), which embed the solute in a medium to capture electrostatic and shift transitions by up to 1 eV in polar solvents. However, PCM overlooks specific solute-solvent , necessitating hybrid approaches for quantitative accuracy in aqueous environments.

Applications

In Chemistry and Materials Science

In chemistry, spectroscopy serves as a cornerstone for qualitative and of molecular compositions and structures. (IR) spectroscopy, particularly Fourier transform infrared (FTIR), enables the identification of by detecting characteristic vibrational bands associated with functional groups, such as C-H stretches in polyolefins or C=O absorptions in polyesters, allowing differentiation among diverse polymer types with high specificity. Similarly, (NMR) spectroscopy is indispensable for elucidating the structures of organic compounds, where chemical shifts, coupling constants, and integration of proton or carbon signals reveal connectivity, , and substitution patterns, as demonstrated in routine applications for synthesizing complex pharmaceuticals. In , spectroscopic techniques provide detailed insights into surface and bulk properties. , through mapping of the G-band and 2D-band intensities, quantifies the number of layers in materials, with the 2D/G intensity ratio serving as a reliable indicator for distinguishing from multilayer structures, essential for optimizing electronic device performance. () characterizes alloy surfaces by analyzing binding energies of core-level electrons, revealing elemental composition, oxidation states, and depth profiles up to 10 , which is critical for understanding corrosion resistance in alloys like . Spectroscopy also facilitates real-time monitoring of chemical reactions, enhancing process optimization. In-situ ultraviolet-visible (UV-Vis) spectroscopy tracks reaction kinetics by observing absorbance changes at specific wavelengths corresponding to reactant consumption or product formation, enabling the determination of rate constants in catalytic processes such as olefin polymerization. Yields in these reactions are quantified by integrating peak areas in spectra, normalized against internal standards, to calculate molar concentrations and conversion efficiencies, as commonly applied in NMR or UV-Vis analyses of organic syntheses.

In Physics and Astronomy

In physics and astronomy, spectroscopy provides essential insights into fundamental physical processes and the properties of celestial bodies by examining the emission, absorption, and scattering of electromagnetic radiation. This technique reveals details about motion, energy distributions, densities, and compositions in environments ranging from laboratory plasmas to distant galaxies, enabling tests of theoretical models and the exploration of universal scales. In astrophysics, spectroscopy exploits the Doppler shift to quantify radial velocities of astronomical objects. The redshift parameter z = \frac{\Delta \lambda}{\lambda}, where \Delta \lambda is the observed wavelength shift from the rest wavelength \lambda, approximates the radial velocity v_r via z \approx v_r / c for velocities much less than the speed of light c. This method detects orbital motions in binary stars and exoplanet systems through periodic line shifts in high-resolution spectra, achieving precisions down to meters per second. Spectral line ratios further enable abundance determinations; for example, the intensity ratio of iron lines to hydrogen lines yields the metallicity [Fe/H], tracing nucleosynthetic processes in stellar atmospheres. Gamma-ray spectroscopy in characterizes radioactive decays and nuclear transitions by resolving discrete energies from excited states. Detectors like high-purity crystals measure gamma-ray intensities to compute branching ratios, defined as the fraction of decays proceeding via a specific , such as gamma emission versus release. In the deuterium-tritium reaction, this approach has quantified the gamma-to- branching ratio at approximately 10^{-5}, informing models of production in controlled . Such measurements validate quantum mechanical predictions and constrain parameters in beyond-Standard-Model theories. For plasma diagnostics, Stark broadening of atomic lines serves as a non-intrusive probe of in ionized gases. The broadening arises from quadratic Stark shifts induced by microelectric fields from nearby electrons and ions, with the line width \Delta \lambda scaling as n_e \sim (\Delta \lambda)^{3/2} for Balmer lines in weakly coupled s. This relation, calibrated against Griem's theory, has diagnosed densities up to $10^{17} cm^{-3} in fusion-relevant edge s. The method complements by providing spatially resolved profiles in high-temperature environments. Cosmological spectroscopy has established the cosmic microwave background (CMB) as a near-perfect blackbody spectrum with temperature T = 2.725 K, as precisely fitted by Planck mission data across microwave frequencies. This thermal relic from recombination at z \approx 1100 constrains and Hubble parameter within Lambda-CDM models. Redshift surveys, employing multi-object spectrographs on telescopes like the , measure galaxy to map large-scale structure, revealing voids, filaments, and the accelerated expansion via . These surveys span billions of light-years, yielding cosmological distances and constraints from over a million spectroscopic .

In Biology and Medicine

Spectroscopy plays a pivotal role in by enabling non-invasive imaging, molecular analysis, and diagnostic assessments of living systems. Techniques such as and (NMR) variants provide insights into biomolecular dynamics and tissue properties, facilitating advancements in disease detection and therapeutic development. In bioimaging, is widely used to study pathways, where single-molecule resonance (smFRET) reveals conformational changes by measuring distances between donor and acceptor fluorophores attached to the protein. The FRET efficiency E quantifies and is given by E = \frac{1}{1 + \left( \frac{R}{R_0} \right)^6 }, where R is the distance between fluorophores and R_0 is the Förster radius characteristic of the donor-acceptor pair, allowing researchers to map folding landscapes in proteins like with high spatiotemporal resolution. This approach has characterized metastable states in multi-domain proteins, aiding understanding of misfolding-related diseases such as Alzheimer's. For medical diagnostics, (MRI), a spatial variant of NMR spectroscopy, exploits differences in proton relaxation times to generate tissue contrast, enabling visualization of soft tissues and abnormalities like tumors without . (PET), which detects gamma rays from positron-emitting radionuclides, complements MRI by quantifying metabolic activity in tissues, such as glucose uptake in cancer cells using tracers like 18F-fluorodeoxyglucose. These methods together enhance diagnostic accuracy, as seen in where PET identifies metastatic sites with sensitivity exceeding 90% in many cases. In , (CD) spectroscopy assesses the of biomolecules and pharmaceuticals by measuring differential absorption of left- and right-circularly polarized light, crucial for verifying enantiomeric purity in that can differ in efficacy and toxicity. Binding affinities are determined via spectroscopic shifts, particularly NMR chemical shift perturbations, where ligand binding induces measurable changes in protein frequencies, allowing quantification of constants (Kd) in the nanomolar to micromolar for lead optimization. This has accelerated fragment-based by providing rapid structural insights into protein-ligand interactions. Emerging applications include hyperspectral endoscopy for cancer detection, which captures reflectance spectra across hundreds of wavelengths to differentiate malignant from healthy tissues based on biochemical composition, achieving detection accuracy of 79% (sensitivity 72%, specificity 84%) in a 2023-2024 study of head and neck cancers.

Historical Development

Early Discoveries

The foundations of spectroscopy were laid in the through Isaac Newton's experiments with , which demonstrated the dispersion of white light into its constituent colors. In 1666, Newton passed sunlight through a prism and observed that white light was not homogeneous but composed of a of colors ranging from to violet, with each color corresponding to rays of different refrangibility. He further confirmed this in his "Experimentum Crucis" by recombining the dispersed rays with a second prism and , showing that the colors were inherent properties of light rather than modifications produced by the prism itself. These observations established that white light is a mixture of all spectral colors in balanced proportions, providing the first empirical insight into the composition of light. In the early 19th century, the known spectrum was extended beyond the visible range. In 1800, British astronomer discovered radiation by placing a thermometer beyond the red end of the solar spectrum dispersed by a and observing elevated temperatures, indicating invisible "heat rays." The next year, identified ultraviolet radiation through its ability to blacken paper more strongly beyond the violet end. These findings broadened the scope of spectroscopic inquiry. Advancements in the 19th century built on these ideas with the invention of more precise instruments and the discovery of spectral lines. In 1814, German physicist Joseph von Fraunhofer constructed an early spectroscope using a slit, collimating lens, prism, and telescope, which allowed him to examine sunlight dispersed into a spectrum. Through this device, Fraunhofer identified approximately 574 dark absorption lines—now known as Fraunhofer lines—superimposed on the continuous solar spectrum, marking the first systematic observation of such features and enabling wavelength measurements. These lines hinted at the interaction between light and matter, though their origins remained unexplained at the time. Further progress came in 1859–1860 when and developed laws relating and , using improved spectroscopes with adjustable slits and prisms to analyze flames and vapors. Their experiments showed that a cool gas absorbs light at the same wavelengths it emits when hot, leading to Kirchhoff's laws: a luminous solid or dense gas produces a continuous ; a luminous gas produces bright lines; and a non-luminous gas interposed between a continuous source and observer produces dark lines at those wavelengths. This work not only enabled chemical analysis through spectral signatures but also integrated diffraction gratings—ruled wires or surfaces that Fraunhofer had pioneered earlier—for higher-resolution dispersion. The principles established by Kirchhoff and Bunsen were soon applied to astronomical observations. In 1868, during a total , French astronomer and English astronomer independently detected a bright yellow emission line at 587.6 nm in the Sun's , which did not match any known . They proposed this indicated a new , named (from the Greek "helios" for sun), marking the first discovery of an element through spectroscopy. Helium was later isolated on in 1895. By the late 19th century, empirical formulas began to describe atomic spectra, culminating in studies of . In 1885, Swiss mathematician Johann Balmer analyzed the visible spectral lines of and proposed an empirical relation fitting their wavelengths, later expressed in terms of wavenumbers for the . This series corresponds to transitions to the n=2 energy level, with lines such as H-alpha at 656 nm. In 1888, Swedish physicist generalized Balmer's formula to other series and elements, introducing the Rydberg equation: \frac{1}{\lambda} = R \left( \frac{1}{n_1^2} - \frac{1}{n_2^2} \right) where \lambda is the wavelength, R is the Rydberg constant (approximately 1.097 × 10^7 m^{-1}), n_1 and n_2 are positive integers with n_1 < n_2. For the Balmer series, n_1 = 2. This formula unified observations of spectral lines across atoms, laying groundwork for understanding atomic structure through radiation-matter interactions. Early spectroscopes, refined with narrow slits for resolution and gratings for even dispersion, were essential to these measurements.

Modern Evolution and Key Milestones

The integration of into spectroscopy marked a pivotal shift in the early , beginning with Niels Bohr's 1913 model of the , which successfully explained the discrete emission lines observed in hydrogen spectra by quantizing electron orbits and linking energy transitions to specific wavelengths. This model laid the groundwork for understanding atomic structure through , though it was limited to hydrogen-like atoms. Building on this, and proposed electron spin in 1925, introducing an intrinsic that accounted for the splitting in atomic spectra, such as the anomalous , thereby refining quantum mechanical interpretations of spectroscopic data. These advancements transitioned spectroscopy from empirical observations to a theoretically driven field. A major breakthrough in (NMR) spectroscopy occurred in 1946 when and Edward Purcell independently developed techniques to measure nuclear magnetic moments in solids and liquids, enabling high-resolution studies of molecular structures. Their work, recognized with the , transformed NMR into a cornerstone of analytical spectroscopy by providing precise information on chemical environments through spin interactions. The invention of the in 1960 by revolutionized spectroscopy by delivering coherent, monochromatic, and intense light sources, enabling unprecedented precision in excitation and detection across spectral regions. This "laser revolution" facilitated nonlinear and time-resolved techniques, dramatically improving signal-to-noise ratios and allowing interrogation of transient species. In the 1980s, Ahmed Zewail advanced this further with laser pulses, earning the 1999 for , which captured ultrafast bond dynamics in real time, revealing reaction mechanisms at atomic scales. Instrumentation evolved significantly in the 1960s with the advent of (FT) spectroscopy, pioneered by developments in and , which allowed multiplexed data collection for faster, higher-resolution and far-infrared spectra compared to dispersive methods. By the 1980s, (CCD) detectors supplanted tubes, offering array-based imaging with quantum efficiencies exceeding 80% and enabling simultaneous multi-wavelength detection in astronomical and . From the 1970s onward, dedicated sources, such as the National Synchrotron Light Source, provided tunable, high-brilliance beams, vastly enhancing absorption and photoelectron spectroscopies for probing electronic structures in materials and biomolecules. The year 2001 marked a milestone with the first generation and characterization of isolated pulses by Hentschel et al., enabling spectroscopy of electron dynamics on sub-femtosecond timescales and opening attosecond science. Post-2015, (AI) has accelerated spectroscopic analysis through models that automate peak identification, noise reduction, and pattern recognition in complex datasets from Raman and spectroscopies, improving accuracy and throughput in high-dimensional data processing.

Educational and Practical Aspects

DIY Spectroscopy Techniques

DIY spectroscopy techniques enable amateurs, students, and educators to explore light-matter interactions using everyday materials and minimal equipment, fostering hands-on understanding of and spectra without specialized access. These methods typically rely on basic optical principles, such as and simple detection, to separate wavelengths of light and observe spectral patterns. Common setups achieve qualitative observations of spectral lines or bands, suitable for educational demonstrations of atomic and molecular behaviors. One accessible setup involves repurposing a (CD) or digital versatile disc (DVD) as a to observe emission spectra. The closely spaced tracks on a CD, with approximately 625 lines per millimeter, or a DVD with about 1350 lines per millimeter, diffract incoming into its component when mounted in a simple enclosure like a cereal box or paper tube with a narrow slit. Light from a source enters through the slit, reflects off the disc's reflective side at an angle, and projects a dispersed onto a viewing surface or camera for capture. This design allows visualization of continuous spectra from white light sources or discrete lines from gaseous emissions, such as those from fluorescent lamps. For enhanced portability, smartphone spectrometer apps or attachments integrate the phone's camera with a small or slit, processing images to display spectra in real-time; for instance, apps like those developed for educational purposes can analyze light from LEDs or flames directly. These tools, often built with 3D-printed holders or household items, provide a digital interface for plotting intensity versus wavelength. Practical experiments using these setups include flame tests to identify metal ions through their characteristic emission lines. In a basic flame test, small amounts of metal salts—such as copper(II) chloride for green emission or strontium chloride for red—are introduced to a Bunsen burner or alcohol-soaked wick flame, exciting electrons to higher energy levels and producing visible colors upon relaxation. Observing the flame through the DIY spectroscope reveals distinct spectral lines, linking the observed hues to specific wavelengths, like the green from copper around 500-570 nm. Another experiment demonstrates absorption spectroscopy by passing white light through household dyes, such as food coloring diluted in water, and viewing the transmitted spectrum. Solutions of blue or red dyes absorb complementary wavelengths, creating dark bands in the otherwise continuous spectrum; for example, tartrazine (yellow food dye) shows strong absorption near 430 nm, observable as a gap in the blue-violet region when analyzed with a smartphone-based setup. The resolution of these basic DIY designs is limited, typically around 5-10 for slit widths of 0.1-0.2 , due to the grating's line density and optical imperfections, preventing separation of closely spaced lines like mercury's yellow pair at 577 and 579 . Calibration improves accuracy by aligning the observed to known ; a common reference is the prominent green mercury emission line at 546.1 from a fluorescent or (CFL), which serves as a fixed point to scale the wavelength axis across the visible range (400-700 ). Multiple lines, such as mercury's blue at 436 , allow for broader . Safety is paramount in DIY spectroscopy to prevent from optical or . When using lasers for alignment or as monochromatic sources, direct eye exposure must be avoided, as even low-power Class 2 lasers can cause damage; indirect viewing through the spectroscope or diffusers is recommended. For flame tests involving metal salts, handle chemicals with gloves to avoid skin contact, work in well-ventilated areas to minimize of fumes, and extinguish s promptly to prevent burns or fires. Protective should be worn during all experiments, and is advised for younger participants to ensure proper handling of heat sources and sharp edges from constructing devices.

Instrumentation and Measurement

Spectroscopic encompasses a range of professional-grade devices designed to precisely measure the between and across various . Central to these systems are components that disperse into its components and detect the resulting signals with high sensitivity. Monochromators and detectors form the core of dispersive spectrometers, while interferometric setups like those in infrared (FT-IR) spectroscopy offer advantages in multiplexed detection. protocols ensure accuracy in and intensity measurements, and techniques such as lock-in amplification address common error sources to enhance signal quality. Monochromators are essential for isolating specific wavelengths in dispersive spectroscopy, typically employing to achieve . A consists of a reflective surface with evenly spaced grooves, where the governs the angular separation of wavelengths:
d \sin \theta = m \lambda
Here, d is the groove spacing, \theta is the , m is the (an integer), and \lambda is the . This describes how incident is diffracted into orders, allowing separation of based on their wavelength-dependent angles; for , the simplifies to highlight the linear with groove density. Blazed gratings optimize by aligning groove facets to reflect constructively into a specific , minimizing losses in applications like UV-visible spectroscopy.
Detectors convert photonic signals into measurable electrical outputs, with photomultiplier tubes (PMTs) and array detectors being prominent choices. PMTs amplify photoelectrons through a series of dynodes, achieving gains up to $10^6 or more, which enables single-photon detection with low noise, ideal for low-light or . In contrast, (CCD) arrays consist of pixelated silicon sensors that capture spatial distributions of light simultaneously, offering high (up to 90%) across a broad spectral range and enabling parallel readout for full-spectrum imaging in array-based systems. PMTs excel in sequential, high-sensitivity measurements, while CCD arrays provide faster acquisition for multichannel data but may introduce read noise in low-signal scenarios. Spectrometers vary in design to balance , speed, and throughput. Scanning spectrometers employ a movable to sequentially pass wavelengths through a single detector like a , providing high (down to 0.1 nm) but requiring longer acquisition times for full . Array spectrometers, conversely, use fixed gratings with detector arrays (e.g., CCDs or photodiode arrays) to disperse and detect the entire simultaneously, enabling rapid measurements (milliseconds per scan) at the cost of slightly lower per-channel . In FT-IR spectroscopy, interferometers replace dispersive elements; a splits a source beam, creates a path-length difference via a moving mirror, and recombines it to produce an interferogram, which is Fourier-transformed into a . This design yields the multiplex (Fellgett) advantage, where all wavelengths contribute to the (SNR) concurrently, improving SNR by a factor of \sqrt{N} (N being the number of resolution elements) compared to scanning dispersive systems under detector-noise-limited conditions./Spectroscopy/Fundamentals_of_Spectroscopy/The_Power_of_the_Fourier_Transform_for_Spectroscopists) Calibration maintains to international standards, beginning with verification using sources like the helium-neon (He-Ne) . Unstabilized He-Ne lasers emit at a vacuum of 632.9908 with a relative standard of $1.5 \times 10^{-6}, serving as a secondary standard for aligning positions or CCD pixel assignments in UV-visible and near-IR spectrometers. Intensity (radiometric) calibration employs blackbody sources, which approximate ideal Planckian radiators to provide known spectral irradiance across temperatures (e.g., 1000–3000 K); by comparing measured spectra to theoretical blackbody curves, instrument response functions are derived, ensuring absolute intensity scales for . Error sources, such as , compromise measurement fidelity by introducing non-dispersed radiation that bypasses the , typically via from optical components or imperfections. Stray light manifests as a constant offset in , causing absorbance errors that deviate from Beer's law, particularly at extremes where it can exceed 0.5% of total and inflate apparent transmittance by up to 10% in UV regions. To mitigate and enhance weak signals, lock-in amplifiers perform phase-sensitive detection: the input signal is multiplied by a reference at the modulation frequency (e.g., from a chopped source), followed by low-pass filtering, which rejects broadband while preserving the correlated signal, achieving dynamic reserves over 120 dB in spectroscopic applications like photoacoustic or modulated measurements.

References

  1. [1]
    Spectroscopy of rocks and minerals and principles of spectroscopy
    Spectroscopy is the study of light as a function of wavelength that has been emitted, reflected or scattered from a solid, liquid, or gas.
  2. [2]
    Types of Spectra and Spectroscopy - NASA Science
    Jul 11, 2018 · The basic premise of spectroscopy is that different materials emit and interact with different wavelengths (colors) of light in different.
  3. [3]
    The Era of Classical Spectroscopy - MIT
    Newton' s analysis of light was the beginning of the science of spectroscopy. It gradually became clear that the sun's radiation has components outside the ...
  4. [4]
    The Era of Modern Spectroscopy - MIT
    The era of modem spectroscopy began with the invention of the laser, which provides intense, collimated monochromatic radiation throughout optical spectral ...
  5. [5]
    Spectroscopy - Center for Astrophysics | Harvard & Smithsonian
    Spectroscopy analyzes light broken down into its spectrum, revealing information about temperature, atoms, and speed of astronomical objects. It uses ...Missing: fundamentals | Show results with:fundamentals
  6. [6]
    Spectroscopy and Principles of Spectroscopy | U.S. Geological Survey
    Extensive tutorial on reflectance spectroscopy, the causes of absorption bands, the absorption and scattering processes that occur when light encounters a ...
  7. [7]
    Definition of Spectroscopy Introduction - Chemicool
    Spectroscopy is the use of the absorption, emission, or scattering of electromagnetic radiation by matter to qualitatively or quantitatively study the matter.
  8. [8]
    Spectroscopy - an overview | ScienceDirect Topics
    Spectroscopy refers to the study of, and the practical methods for examining, the interaction of electromagnetic radiation (light) and matter.<|separator|>
  9. [9]
  10. [10]
    What Is Spectroscopy? - SGS PSI - Polymer Solutions
    Mar 27, 2014 · Later it was extended to cover all frequencies of electromagnetic radiation, from long radio waves to short gamma rays, i.e., the ...Missing: scope | Show results with:scope
  11. [11]
    Spectra and What They Can Tell Us
    Spectra can be produced for any energy of light, from low-energy radio waves to very high-energy gamma rays. Each spectrum holds a wide variety of information.
  12. [12]
    Spectrophotometry vs. Spectroscopy - HunterLab
    Apr 11, 2024 · Spectroscopy observes how radiated matter and energy interact, while spectrophotometry measures light absorption in a chemical substance. INFO ...
  13. [13]
    Spectroscopy vs. Spectrometry in OES - Verichek Technical Services
    Essentially, spectroscopy provides the theoretical framework for understanding energy-matter interactions, while spectrometry offers the practical tools for ...
  14. [14]
    4.2: Understanding Atomic Spectra - Chemistry LibreTexts
    Jun 26, 2023 · A continuous spectrum is a smooth color gradient. Emission lines show the exact wavelengths of the electrons that are emitted.Continuous Spectrum · Atomic Emission Spectra · Atomic Absorption Spectra
  15. [15]
    [PDF] Astronomical Spectroscopy - Lowell Observatory
    The spectral resolution is characterized as R = λ/∆λ, where ∆λ is the resolution ele- ment, the difference in wavelength between two equally strong ( ...
  16. [16]
    The Importance of Dynamic Range and Signal to Noise Ratio in ...
    Dynamic range is the ratio of maximum to minimum signal intensities. Signal to noise ratio (SNR) is signal intensity divided by noise intensity.
  17. [17]
    Signal-to-noise ratio in spectrometry - Silicann Systems
    Jun 24, 2025 · Signal-to-noise ratio (SNR) measures the difference between the desired useful signal and the unwanted background noise of a sensor.
  18. [18]
    Spectral Resolution Amelioration by Deconvolution (SPREAD ... - NIH
    Spectral resolution defines the ability to distinguish two closely spaced peaks in a spectrum. It is one of the most important determinants of the quality of ...
  19. [19]
    6.3 How is energy related to the wavelength of radiation?
    The energy associated with a single photon is given by E = h ν , where E is the energy (SI units of J), h is Planck's constant (h = 6.626 x 10–34 J s), and ν is ...
  20. [20]
    CHEM 101 - Electromagnetic radiation and quantum theory
    Oct 10, 2023 · The Planck relationship states that the energy of a photon of electromagnetic radiation is proportional to its frequency, and inversely ...
  21. [21]
    spectroscopy ch 3
    Atomic absorption and emission spectroscopy in the ultraviolet and visible regions are used to determine metals in samples derived from air, water or solids.
  22. [22]
    Raman Spectroscopy | Scientific Technical Services - WWU's SciTech
    Raman spectroscopy uses the inelastic scattering of light (Raman scattering) to measure the vibrational modes of molecules.
  23. [23]
    [PDF] Light Absorption (and Optical Losses) - MIT OpenCourseWare
    Medium. I = I o. ∙e. -α∙l. Position. Simple Derivation of Beer-Lambert's Law: dI z = -σ ∙ N ∙ dz. I z ln I z. )+ C. ( )= -(σ ∙ N ∙ z ln I. 0. ( )= -(σ ∙ N ∙0. (. ) ...
  24. [24]
    [PDF] Selection Rules: - MSU chemistry
    associated selection rules. ▫ Selection rules originate from the quantum mechanical description of electromagnetic radiation interaction with matter.
  25. [25]
    Quantisierung als Eigenwertproblem - Wiley Online Library
    Quantisierung als Eigenwertproblem. E. Schrödinger,. E. Schrödinger. Zürich ... Download PDF. back. Additional links. About Wiley Online Library. Privacy ...
  26. [26]
    [PDF] Absorption, Emission and Fluorescence Spectroscopies
    Emission is the process that creates a photon and takes the the atom or molecule in an excited state back to the ground state. ... P = power of transmitted light ...
  27. [27]
    Raman Scattering - HyperPhysics
    Such inelastic scattering is called Raman scattering. Like Rayleigh scattering, the Raman scattering depends upon the polarizability of the molecules. For ...
  28. [28]
    Brillouin Scattering Theory
    Brillouin light scattering is the result of the interaction between a photon of light and the acoustic phonons of a material.
  29. [29]
    application to phase sensitive spectroscopy and near-field nanoscopy
    The general principle consists in modulating the phase in the reference arm of an interferometer by adding a variable delay length and measuring the resulting ...<|separator|>
  30. [30]
    Optimum Sample Thickness for Trace Analyte Detection with Field ...
    Broadband IR spectroscopy, on the other hand, profits from large interaction cross sections, potentially affording a unique combination of detection sensitivity ...Scaling Laws And Optimum... · Experimental Setup · Results
  31. [31]
    Limits of Detection in Spectroscopy (PDF)
    Dec 1, 2003 · A diagrammatic view of the concept of detection limits is presented that effectively shows its relationship to the reproducibility of measurements on the blank.
  32. [32]
    A Primer on Quantum Numbers and Spectroscopic Notation
    Quantum numbers are n (principal), l (azimuthal), m (magnetic), and s (spin). Spectroscopic notation describes the state of electrons in an atom or ion.
  33. [33]
    [PDF] Zeeman Effect in Mercury - University of Washington
    This scheme of coupling angular momenta for more than one electron is called LS coupling, and is also known as Russell-Saunders coupling. (The common short ...
  34. [34]
    Principles and Applications of Atomic Absorption Spectroscopy
    The chapter discusses atomic absorption spectroscopy, and its principles and applications. The term atomic absorption spectroscopy was familiar to only a ...
  35. [35]
    1.4: Introduction to Atomic Absorption Spectroscopy
    Aug 28, 2022 · There are many applications of atomic absorption spectroscopy (AAS) due to its specificity. These can be divided into the broad categories ...Brief overview of atomic... · Applications of Atomic... · Obtaining Measurements<|separator|>
  36. [36]
    Flame Photometric Determination of Alkali and Alkaline Earth ...
    The flame photometric determination of calcium in phosphate, carbonate, and silicate rocks. Analytica Chimica Acta 1957, 17 , 521-525.
  37. [37]
    Flame Photometry (Theory) : Inorganic Chemistry Virtual Lab
    Flame photometry, a branch of atomic spectroscopy, determines metal ion concentrations by measuring emitted light from excited atoms in a flame.
  38. [38]
    [PDF] Multi-Elemental Analysis of Aqueous Geological Samples by ...
    This report details multi-elemental analysis of aqueous geological samples using Inductively Coupled Plasma-Optical Emission Spectrometry, by Todor I. Todorov, ...
  39. [39]
    [PDF] EPA Method 6010C (SW-846): Inductively Coupled Plasma
    2.2. This method describes multielemental determinations by ICP-AES using sequential or simultaneous optical systems and axial or radial viewing of the plasma.<|separator|>
  40. [40]
    The solar spectrum from Fraunhofer to Skylab—an appreciation of ...
    The history of spectroscopy is reviewed with particular reference to the observation and identification of the Fraunhofer lines in the solar spectrum.<|control11|><|separator|>
  41. [41]
  42. [42]
    [PDF] Lecture 2: Rotational and Vibrational Spectra
    Vibration-Rotation Spectra (IR). Vibration-Rotation spectrum of CO. (from FTIR) ... Not IR-active, use Raman spectroscopy! ▫. ← for homonuclear molecules.
  43. [43]
    [PDF] Harmonic and Anharmonic Oscillator Models for Pure Vibrational ...
    Aug 2, 2021 · The equation (14) shows that the energy levels of anharmonic oscillator are not equidistant, but their separation decreases slowly with ...Missing: hν | Show results with:hν
  44. [44]
    [PDF] fifty years of the jabłoński diagram - Chemistry
    Jabłoński missed the idea of the intersystem crossing ratio [13] as an intrinsic molecular property which would permit a variation of M-state yields ...<|separator|>
  45. [45]
    Latest FT-IR Spectroscopy Research Applications
    Sep 3, 2024 · Key applications include hydrogen bonding studies, environmental monitoring, food analysis, and clinical diagnostics. The article underscores FT ...
  46. [46]
    A Comprehensive Review of Raman Spectroscopy in Biological ...
    Dec 3, 2024 · Raman spectroscopy has been proven to be a fast, convenient, and nondestructive technique for advancing our understanding of biological systems.Principles and Techniques of... · Applications of Raman... · Challenges in Raman...
  47. [47]
    Raman Spectroscopy for In-Line Water Quality Monitoring
    Since water is a weak Raman scatterer, Raman spectroscopy is superior to other vibrational spectroscopies particularly in some applications such as biomedical ...
  48. [48]
    [PDF] "Mass and Isotope-selective Infrared Spectroscopy" in
    shift of −4.1% will occur due to the change in the reduced mass. The overtone vibrations of the CH-chromophore will probably resemble local diatomic CH ...
  49. [49]
    Collective dynamics in the condensed phase - Baldini Lab
    Collective dynamics involve cooperative, wave-like motion of many particles, such as phonons, magnons, and excitons, in condensed matter.
  50. [50]
    Direct and Indirect Band Gap Semiconductors - DoITPoMS
    In a direct band gap semiconductor, the top of the valence band and the bottom of the conduction band occur at the same value of momentum.Missing: theory seminal paper
  51. [51]
    Direct measurement of key exciton properties: Energy, dynamics ...
    Jun 7, 2021 · Excitons, Coulomb-bound electron–hole pairs, are the fundamental excitations governing the optoelectronic properties of semiconductors.<|separator|>
  52. [52]
    Excitons and excitonic materials | MRS Bulletin
    Sep 1, 2024 · An exciton is a bound pair of negatively charged electron and positively charged hole (electron vacancy within a solid), both of which are held together by ...
  53. [53]
    X-ray photoelectron spectroscopy: Towards reliable binding energy ...
    An essential part of XPS analysis is the evaluation of elemental composition in the surface region. ... surface might shift the binding energy of the carbon.
  54. [54]
    Chemical significance of x-ray photoelectron spectroscopy binding ...
    Oct 6, 2023 · The principal intent of this Perspective is to review the mechanisms that are responsible for the shifts of binding energies, ΔBE, observed in x-ray ...
  55. [55]
    Ultraviolet photoelectron spectroscopy: Practical aspects and best ...
    UPS measures valence band energies of surfaces using UV light, measuring ionization energies of valence shell electrons, and work function.
  56. [56]
    Ultraviolet Photoelectron Spectroscopy of Solids - SpringerLink
    As we shall see, its main strength lies in its unique ability to explore the electronic structure in the conduction/valence band region of a wide variety of ...
  57. [57]
    Raman Spectroscopy and Imaging of Low Energy Phonons
    Sep 1, 2015 · Raman spectroscopy of solid-state materials involves the inelastic scattering of light by phonons, quanta that have the energy of lattice ...
  58. [58]
    Calculated phonon modes, infrared, and Raman spectra in ZnGeGa ...
    Aug 19, 2020 · Here, we investigate the vibrational properties of this new compound and provide predictions for its related infrared and Raman spectra, which ...INTRODUCTION · Infared spectra · Raman spectra · Phonon dispersion and...
  59. [59]
    5.5: Vibrational States and Phonon Dispersion Curves
    Apr 7, 2024 · Phonons are vibrational modes of a crystal that form wave packets characterized by a frequency ω and momentum ℏ ⁢ k . In a solid, they can be ...
  60. [60]
    [PDF] 4. Phonons - DAMTP
    ω(k) k. Figure 56: Phonon dispersion relation for a monatomic chain. Many ... of k, the dispersion relation for phonons is linear ! ⇡ rm ak. This is in ...
  61. [61]
    Defect Absorption - an overview | ScienceDirect Topics
    The transitions involve localized gap states which result from various lattice defects, mostly intrinsic ones such as under- and over-coordinated lattice sites, ...
  62. [62]
    Transition metal impurities in silicon: computational search ... - Nature
    Aug 19, 2022 · Localized spin defect states are required for optically active spin defects ... states and the mid-gap defect states. Also, the absorption ...
  63. [63]
    Nuclear quadrupole spectroscopy (NQR) - Spect homepage
    Figure 1: Nuclear quadrupole resonance requires that the nuclei under scrutiny display electric quadrupole moments.[2] electric quadrupole. When a nucleus ...
  64. [64]
    57Fe Mössbauer Spectroscopy as a Tool for Study of Spin States ...
    Feb 18, 2021 · For example, the isomer shifts typical of LS FeII compounds are δ = −0.1 to 0.2 mm/s, the quadrupole splitting is Δ = 0.2 to 1.9 mm/s, and ...
  65. [65]
    [PDF] Introduction to NMR
    The fundamental relationship is. µ=γS γ (gamma) is the gyromagnetic ratio. It depends on the nucleus. For the proton it is. 2,675*108 rad/s/T.
  66. [66]
    Chemical shift referencing
    The chemical shift is calculated as follows. δ = (ν - ν0)/ν0. = (400.13286373 - 400.13000000)/ 400.13000000. = 0.000007157 = 7.157 ppm. For other nuclei we use ...
  67. [67]
    Solid state nuclear magnetic resonance of polymers - ScienceDirect
    This work aims to present a review of applications of SSNMR to study different polymer systems, carried out at the NMR group in Córdoba, Argentina from the year ...
  68. [68]
    4.2: Hyperfine Hamiltonian - Chemistry LibreTexts
    Mar 24, 2022 · To first order, the contribution of the hyperfine interaction to the energy levels is then given by m S ⁢ m I ⁢ A . In the EPR spectrum, each ...
  69. [69]
    4.8: EPR Spectroscopy - Chemistry LibreTexts
    Aug 28, 2022 · Electron paramagnetic resonance spectroscopy (EPR) is a powerful tool for investigating paramagnetic species, including organic radicals, inorganic radicals, ...
  70. [70]
    Breaking Barriers in Ultrafast Spectroscopy and Imaging Using 100 ...
    Jul 10, 2023 · In the pump–probe technique, or “transient absorption” (TA, Figure 2 ... Typically, two probe pulses are used per pump pulse, optically ...
  71. [71]
    High‐Resolution Time‐Correlated Single‐Photon Counting Using ...
    Jul 20, 2022 · In fluorescence lifetime techniques, the distribution of time-delays between an excitation laser pulse and the emitted fluorescence photons ...
  72. [72]
    Review Time-dependent density-functional theory for molecules and ...
    Nov 30, 2009 · This article helps to fill the void by first giving a historical review of TDDFT, with emphasis on molecular excitations and aspects of TDDFT ...
  73. [73]
    Single‐reference coupled cluster methods for computing excitation ...
    Sep 12, 2019 · This review examines the various approximation schemes with particular emphasis on their performance for excitation energies and summarizes the best state-of- ...2 The Eom--Ccsd Method · 3.1 Perturbative Techniques · 5 Benchmarks<|separator|>
  74. [74]
    Cost-Effective Simulations of Vibrationally-Resolved Absorption ...
    Apr 5, 2023 · Computing UV/vis spectra from the adiabatic and vertical Franck-Condon schemes with the use of Cartesian and internal coordinates. J. Chem ...
  75. [75]
    Efficient anharmonic vibrational spectroscopy for large molecules ...
    Sep 10, 2014 · This article presents a general computational approach for efficient simulations of anharmonic vibrational spectra in chemical systems.<|separator|>
  76. [76]
    Deep Learning Spectroscopy: Neural Networks for Molecular ...
    Jan 29, 2019 · 10 Conclusion. In summary, we demonstrated that deep neural networks can learn spectra to 97% accuracy and peak positions to within 0.19 eV. ...
  77. [77]
    Mind the basis set superposition error - ScienceDirect.com
    Aug 20, 2010 · The quantities calculated (geometry, binding energies, frequencies, etc.) need to be corrected for the basis set superposition error.
  78. [78]
    Polarizable continuum models for quantum-mechanical descriptions
    Apr 22, 2016 · Polarizable continuum solvation models are nowadays the most popular approach to describe solvent effects in the context of quantum mechanical calculations.
  79. [79]
    Full article: Contributions of Fourier transform infrared spectroscopy ...
    Aug 24, 2020 · The aim of this review is to discuss and highlight the recent advances in FTIR (spectroscopy and chemical imaging) techniques that are used to ...
  80. [80]
    Structural Elucidation with NMR Spectroscopy: Practical Strategies ...
    May 14, 2008 · NMR spectroscopic methods for structural elucidation in organic chemistry are reviewed. Detailed procedures and recommended experimental parameters areprovided.Missing: seminal | Show results with:seminal
  81. [81]
    Raman spectroscopy of graphene-based materials and its ...
    Jan 25, 2018 · Raman spectroscopy is a versatile tool to identify and characterize the chemical and physical properties of these materials, both at the laboratory and mass- ...
  82. [82]
    Review on surface-characterization applications of X-ray ...
    This is a pedagogical review on XPS based surface characterization. Applications in materials, nanotechnology and corrosion evaluation are reviewed.
  83. [83]
    In Situ UV–Vis–NIR Absorption Spectroscopy and Catalysis
    Feb 26, 2024 · This review highlights in situ UV–vis–NIR range absorption spectroscopy in catalysis. A variety of experimental techniques identifying reaction mechanisms, ...
  84. [84]
    A Method for Converting HPLC Peak Area from Online Reaction ...
    Jan 10, 2023 · (3) Calculate concentration. Multiply the peak area of each chemical species by the corresponding response factor to calculate concentration.
  85. [85]
    Hyperspectral Imaging for Non-destructive Testing of Composite ...
    Oct 13, 2022 · In this study, it is proposed to identify the constitutive material and classify local defects of composite specimens.
  86. [86]
    Doppler shift - Imagine the Universe! - NASA
    Sep 24, 2020 · This is an example of the Doppler shift, and it is an effect that is associated with any wave phenomena (such as sound waves or light).Missing: spectroscopy | Show results with:spectroscopy
  87. [87]
    Radial Velocity Planet Resources in the Exoplanet Archive
    Apr 27, 2021 · The motion is detected using high-resolution spectroscopy to measure tiny Doppler shifts in the stellar spectrum toward the blue or red as ...
  88. [88]
    A Concise Treatise on Converting Stellar Mass Fractions to ...
    Nov 18, 2022 · Astronomers define the abundance of an element starting from a ratio with respect to the number of hydrogen atoms fixed at a trillion (or 1012 ...
  89. [89]
    Gamma-ray Spectroscopy in Low-Power Nuclear Research Reactors
    Jan 26, 2024 · Gamma rays arising from either nuclear reactions or decay ... branching ratios. For example, we found that 140La has photopeaks at ...Gamma-Ray Spectroscopy In... · 2. Materials And Methods · 3. Results<|separator|>
  90. [90]
    Measurement of the Gamma-Ray-to-Neutron Branching Ratio for the ...
    The gamma-ray-to-neutron branching ratio for the deuterium-tritium reaction was determined in magnetic confinement fusion plasmas at the Joint European Torus.
  91. [91]
    Stark broadening for diagnostics of the electron density in non ...
    Apr 29, 2014 · An easy way to determine simultaneously the electron density and temperature in discharges using Stark broadening has been reported,5 but this ...
  92. [92]
    Interpretation of Stark broadening measurements on a spatially ...
    Dec 16, 2022 · In thermal plasma spectroscopy, Stark broadening measurement of hydrogen spectral lines is considered to be a good and reliable measurement for electron ...
  93. [93]
    Planck and the cosmic microwave background - ESA
    Planck's predecessors (NASA's COBE and WMAP missions) measured the temperature of the CMB to be 2.726 Kelvin (approximately -270 degrees Celsius) almost ...
  94. [94]
    Measuring our Universe from Galaxy Redshift Surveys - PMC
    We describe several methodologies of using galaxy redshift surveys as cosmological probes, and then summarize the recent results from the existing surveys.
  95. [95]
    Redshift Surveys and Cosmology - M. Colless
    It is important to note that the redshift sample generated by this spectroscopic survey will be supplemented by photometric redshifts with a precision of ...
  96. [96]
    Fluorescent Probes and Quenchers in Studies of Protein ... - PubMed
    Sep 11, 2023 · Fluorescence spectroscopy provides numerous methodological tools for structural and functional studies of biological macromolecules and ...
  97. [97]
    Single-molecule fluorescence spectroscopy maps the folding ...
    Oct 11, 2011 · ... FRET based method is developed and used to identify six metastable states in the folding landscape of the three-domain protein adenylate kinase.
  98. [98]
    Single-Molecule FRET Spectroscopy and the Polymer Physics of ...
    Jul 5, 2016 · The properties of unfolded proteins have long been of interest because of their importance to the protein folding process.
  99. [99]
    Reliability and accuracy of single-molecule FRET studies for ...
    Mar 27, 2023 · Our work demonstrates that smFRET is able to characterize challenging and realistic protein systems with conformational dynamics on timescales ...
  100. [100]
    Molecular aspects of magnetic resonance imaging and spectroscopy
    Magnetic resonance imaging (MRI) is a well known diagnostic tool in radiology that produces unsurpassed images of the human body, in particular of soft ...
  101. [101]
    Radiochemistry for positron emission tomography - Nature
    Jun 5, 2023 · In this work, we provide an overview of commonly used chemical transformations for the syntheses of PET tracers in all aspects of radiochemistry.
  102. [102]
    Nuclear Magnetic Resonance Spectroscopy in Clinical ... - PubMed
    Sep 20, 2021 · This review focus on the potential of NMR in the context of clinical metabolomics and personalized medicine.
  103. [103]
    Circular dichroism in drug discovery and development - PubMed
    Circular dichroism spectroscopy is the technique of choice for determining the stereochemistry of chiral drugs and proteins.
  104. [104]
    Rapid protein-ligand costructures using chemical shift perturbations
    Jan 16, 2008 · Structure-based drug discovery requires the iterative determination of protein-ligand costructures in order to improve the binding affinity ...Missing: spectroscopy development
  105. [105]
    NMR-driven structure-based drug discovery by unveiling molecular ...
    May 31, 2025 · NMR spectroscopy has become an indispensable tool in structure-based drug design, especially in the context of fragment-based drug design ...
  106. [106]
    Endoscopic In Vivo Hyperspectral Imaging for Head and Neck ...
    Nov 10, 2024 · In this study, we demonstrated that hyperspectral imaging (HSI) is a feasible tool for rapid in vivo visualization of tissue differences during ...Missing: endoscopy | Show results with:endoscopy
  107. [107]
    Integrating AI with Advanced Hyperspectral Imaging for Enhanced ...
    Aug 8, 2025 · This increases superficial tissue contrast, assisting in the detection of lesions through light filtration within narrow bands. Recent ...
  108. [108]
    'A Letter of Mr. Isaac Newton … containing his New Theory about ...
    A Letter from Liege concerning Mr Newton's Experiment of the colour'd Spectrum [Philosophical Transactions 128 (25 September 1676)]. [Diplomatic Text] ...
  109. [109]
    The Science of Color - Smithsonian Libraries
    In the 1660s, English physicist and mathematician Isaac Newton began a series of experiments with sunlight and prisms. He demonstrated that clear white ...
  110. [110]
    Spectrometers for Elemental Spectrochemical Analysis, Part I
    Jan 1, 2010 · The grating was introduced in 1814 by Joseph von Frauenhofer for astronomical studies. In the X-ray region of the electromagnetic (EM) spectrum, ...
  111. [111]
    Joseph von Fraunhofer
    Around 1814, Fraunhofer started to investigate this phenomenon using a spectrometer of his own invention and in the process discovered 574 dark lines among the ...Missing: primary | Show results with:primary
  112. [112]
    Happy Sesquicentennial, Spectroscopy!
    Jun 1, 2010 · In 1860, physicist Gustav Kirchhoff and chemist Robert Bunsen (1) published a long article detailing their investigations with a spectroscope ( ...Missing: original papers
  113. [113]
    Kirchhoff and Bunsen on Spectroscopy - chemteam.info
    One of us in his work "on the relationship between emission and absorption of bodies for heat and light" (Kirchhoff, these Ann. 109, p. 275) has proved ...Missing: laws 1859-1860
  114. [114]
    [PDF] Johann Jacob Balmer (1825-1898) - UB
    Using measurements by H. W. Vogel and by Huggins of the ultraviolet lines of the hydrogen spectrum I have tried to derive a formula which will represent the ...
  115. [115]
    Getting the numbers right - the lonely struggle of Rydberg | Feature
    Johannes Rydberg was one of the grandfathers of modern-day physics and ... Rydberg's new formula was: v = R(1/n12 – 1/n22). Here, v is the wavenumber of ...Missing: 1888 | Show results with:1888
  116. [116]
    6.4 Bohr's Model of the Hydrogen Atom - University Physics Volume 3
    Sep 29, 2016 · ... 1913, was the first quantum model that correctly explained the hydrogen emission spectrum. Bohr's model combines the classical mechanics of ...
  117. [117]
    Electron spin - HyperPhysics Concepts
    In 1925, Samuel A. Goudsmit and George E. Uhlenbeck postulated that the electron had an intrinsic angular momentum, independent of its orbital characteristics.
  118. [118]
    [PDF] THE HISTORY OF THE ELECTRON? - University of Pittsburgh
    Zeeman effect. Uhlenbeck and Goudsmit (1925) accounted for this splitting by positing the internal spin of the electron. While the other shifts in elec ...<|separator|>
  119. [119]
    The Nobel Prize in Physics 1952 - NobelPrize.org
    The Nobel Prize in Physics 1952 was awarded jointly to Felix Bloch and Edward Mills Purcell for their development of new methods for nuclear magnetic precision ...
  120. [120]
    Laser - This Month in Physics History | American Physical Society
    The principle of the laser dates back to 1917, when Albert Einstein first described the theory of stimulated emission.
  121. [121]
    Press release: The 1999 Nobel Prize in Chemistry - NobelPrize.org
    With femtosecond spectroscopy we can for the first time observe in 'slow motion' what happens as the reaction barrier is crossed and hence also understand the ...
  122. [122]
    The History and Current Status of Fourier Transform Spectroscopy
    Jacquinot and J. Connes started work on a Michelson interferometer to be used as a multiplex spectrometer in the near ir and visible. In addition, Fellgett ...
  123. [123]
    Array detectors are transforming optical spectroscopy
    Enthusiastic predictions in the 1980s that charge-coupled device (CCD) and charge-injection device (CID) array detectors would revolutionize optical ...Missing: advancement | Show results with:advancement
  124. [124]
    [PDF] Impact of Synchrotron Radiation on Materials Research
    1960s, synchrotron radiation has vastly enhanced the utility of pre- existing and contemporary tech- niques, such as X-ray diffraction and. X-ray photoelectron ...
  125. [125]
    Attosecond Electron Dynamics in Molecules | Chemical Reviews
    May 10, 2017 · The first characterization of isolated attosecond pulses was performed in 2001 by Hentschel et al. ... attosecond spectroscopy to the nonlinear ...
  126. [126]
    Machine learning for analysis of experimental scattering and ...
    Nov 22, 2023 · In this perspective, we describe the application of supervised and unsupervised ML to experimental scattering and spectroscopy data analysis.
  127. [127]
    [PDF] Spectroscopy and Remote Sensing - SPIE
    Thus, for light normally incident on the grating, the grating equation reduces simply to Equation 5. mλ = d sin θ. (5) ... grating monochromator to analyze ...Missing: explanation | Show results with:explanation
  128. [128]
    Spectrometer with multichannel photon-counting detector for beam ...
    Dec 22, 2015 · Linear CCD arrays with an electronic shutter in principle allow superposing of high speed, low noise, and a high quantum efficiency. The paper4 ...
  129. [129]
    Spark OES Analysis: CMOS or PMT Detectors? - AZoM
    May 24, 2023 · PMTs are highly specialized detectors that can detect the presence of a few specific elements with very low detection limits. In contrast, CCDs ...<|separator|>
  130. [130]
    Comparison between Array and Scanning UV Vis Spectrophotometers
    Scanning spectrophotometer, usually equipped with a Tungsten-Deuterium lamp · Array spectrophotometer, usually equipped with a Xenon lamp.
  131. [131]
    Uncalibrated Helium-Neon Lasers in Length Metrology | NIST
    Sep 1, 2009 · The adopted value for vacuum wavelength of an uncalibrated helium-neon laser is 632.9908 nm, and the relative standard uncertainty is 1.5x10^-6.
  132. [132]
    Blackbody-based calibration for temperature calculations in the ...
    This paper presents, the results of a method used to create a blackbody-based referenced calibration curve for a spectrometer in the visible and near-IR ...
  133. [133]
    Errors in Spectrophotometry and Calibration Procedures to Avoid ...
    If the stray light occurs exclusively at the longwave or shortwave side, like at the ends of the spectral range, the substance may only have a longwave or ...
  134. [134]
    Principles of Lock-in Detection | Zurich Instruments
    Lock-in amplifiers use the knowledge about a signal's time dependence to extract it from a noisy background. A lock-in amplifier performs a multiplication of ...