Fact-checked by Grok 2 weeks ago

Raman scattering

Raman scattering, also known as the Raman effect, is the of photons by atoms or molecules in a , resulting in a shift of the scattered light's relative to the incident light due to with vibrational, rotational, or other low-energy excitations. This phenomenon contrasts with elastic , where the remains unchanged, and only a tiny fraction (approximately 1 in 10^6 to 10^7 photons) of the scattered light exhibits the Raman shift, which provides a spectroscopic fingerprint of the 's molecular structure. The effect was discovered in 1928 by Indian physicist Sir Chandrasekhara Venkata Raman and his student during experiments on the scattering of monochromatic light in liquids such as and , where they observed modified wavelengths in the scattered spectrum. Independently, Soviet physicists Grigory Landsberg and Leonid Mandelstam reported similar observations in crystals that same year. For this groundbreaking work on light scattering, Raman was awarded the 1930 , becoming the first Asian recipient in the sciences. Raman scattering forms the basis of , a non-destructive analytical technique that probes , , polymorphism, and in solids, liquids, and gases without requiring . The technique excels in identifying symmetric molecular vibrations inaccessible to and has applications in fields ranging from and pharmaceuticals to forensics and biomedical imaging, often enhanced by lasers for greater sensitivity since the . Quantum mechanically, it is described by the Kramers-Heisenberg-Dirac dispersion formula, involving virtual energy states and changes in the scatterer.

Introduction and History

Definition and Phenomenon

Raman scattering is the inelastic scattering of photons by molecules, in which the scattered light experiences a shift in frequency due to the exchange of energy with the molecular vibrational or rotational modes. This process occurs when an incident photon interacts with a molecule, temporarily exciting it to a virtual intermediate state before the molecule returns to a different vibrational or rotational energy level, resulting in the emission of a scattered photon with altered energy. Unlike elastic , where the scattered retains the same frequency as the incident light with no net transfer, Raman scattering involves a measurable energy exchange, leading to either Stokes scattering (lower frequency, longer wavelength) or anti-Stokes scattering (higher frequency, shorter wavelength). In Stokes Raman scattering, the gains from the , transitioning to a higher vibrational or rotational state, while in anti-Stokes scattering, the loses to the , starting from an already excited state. The basic process can be schematically described as follows: an incident is absorbed by the , promoting it to a short-lived that does not correspond to a real electronic ; from this state, the emits a scattered while relaxing to a different vibrational or rotational level in the ground electronic state. This mediation distinguishes Raman scattering from resonant absorption-emission processes, as the exists only transiently during the scattering event. Raman signals are inherently weak, typically on the order of 10^{-6} to 10^{-7} of the incident light intensity, due to the low probability of inelastic scattering events compared to elastic ones. The intensity depends on factors such as the laser excitation wavelength—shorter wavelengths generally enhance scattering cross-sections—and the sample's molecular polarizability and concentration. For example, in gaseous samples, rotational Raman scattering is often prominent due to well-defined molecular orientations, whereas in liquids or solids, vibrational modes dominate the spectra because of denser intermolecular interactions and broader rotational broadening.

Discovery and Early Development

The Raman effect was discovered in 1928 by Indian physicist Chandrasekhara Venkata Raman and his student at the Indian Association for the Cultivation of Science in Calcutta. The phenomenon had been theoretically predicted in 1923 by Austrian physicist Adolf Smekal. They observed the inelastic scattering of light while passing through various liquids, such as , , and , using a spectrograph to detect frequency shifts in the scattered light relative to the incident beam. These shifts, later identified as corresponding to molecular vibrational and rotational transitions, were reported in their seminal paper published on 16 February 1928. For this groundbreaking work on light scattering, Raman was awarded the in 1930, becoming the first Asian recipient in the sciences. Independently, Soviet physicists Grigory Landsberg and Leonid Mandelstam observed the same phenomenon just days later, on 21 February 1928, at . Their experiment involved illuminating a quartz with a mercury and recording the scattered at a 90-degree angle using a spectrograph, which revealed modified spectral lines distinct from . This confirmation was published in July 1928, highlighting the effect's occurrence in solids as well as liquids. Early theoretical interpretations of the Raman effect emerged in the late 1920s and 1930s, building on the Kramers-Heisenberg-Dirac dispersion formula. Developed initially by Hendrik Kramers and in 1925 and refined by in 1927, the formula provided a quantum mechanical framework for light scattering by atoms and molecules, treating the process as occurring through virtual intermediate states rather than real electronic transitions. This approach explained the inelastic nature of the scattering without requiring in the intermediate steps, aligning with the observed frequency shifts. In the pre-laser era, experimental challenges dominated Raman studies due to the inherently weak intensity of scattered signals, which were typically a million times fainter than . Long exposure times—often several hours—were required on photographic plates to capture spectra, limiting observations to highly samples like liquids and crystals under intense illumination from mercury arcs. Despite these hurdles, the 1930s saw the development of dedicated Raman spectrometers, such as those pioneered by George Placzek, incorporating and improved spectrographs for better . These advancements enabled initial applications in , where Raman spectra were used to identify molecular structures and vibrational modes in compounds, complementing for structural elucidation in the 1930s and 1940s.

Molecular and Quantum Foundations

Vibrational and Rotational Modes

Molecules possess three degrees of freedom per atom, corresponding to translation along the x, y, and z axes, resulting in a total of 3N degrees of freedom for a molecule with N atoms. Three of these are translational degrees of freedom for the molecule as a whole, and for non-linear molecules, there are three rotational degrees of freedom, leaving 3N-6 vibrational degrees of freedom; for linear molecules, there are only two rotational degrees of freedom, yielding 3N-5 vibrational degrees of freedom. These vibrational modes represent the internal oscillations of atoms relative to one another, which are crucial for Raman scattering as they correspond to the energy changes observed in scattered light. Vibrational motion in molecules is often approximated using the quantum harmonic oscillator model, where the potential energy is given by V = \frac{1}{2} k (x - x_0)^2, with k as the force constant and x_0 as the equilibrium position. In this model, the energy levels are quantized as E_v = \hbar \omega (v + \frac{1}{2}), where v is the vibrational quantum number (an integer starting from 0), \hbar is the reduced Planck's constant, and \omega is the angular vibrational frequency. This approximation assumes a parabolic potential, leading to equally spaced energy levels separated by \hbar \omega, which provides a foundational understanding of the discrete energy transitions involved in Raman processes. Rotational motion is described by the approximation, treating the molecule as a fixed structure rotating about its . The rotational energy levels are given by E_J = B J(J+1), where J is the rotational (a non-negative integer), and B = \frac{\hbar^2}{2I} is the rotational constant, with \hbar as the reduced Planck's constant and I as the . Each level J has a degeneracy of $2J+1, reflecting the possible projections of along a quantization axis, and the energy spacing increases quadratically with J, influencing the in Raman spectra. In polyatomic molecules, the vibrational manifest as normal modes, which are collective oscillations where all atoms move in phase with the same frequency. These modes are classified into stretching vibrations, such as symmetric stretches (where bond lengths change in unison) and asymmetric stretches (where bonds alternately lengthen and shorten), as well as bending modes including scissoring, rocking, wagging, and twisting. The specific forms of these normal modes are determined by the molecule's symmetry, often analyzed using point groups like C_{2v} for or D_{3h} for , which dictate the irreducible representations and degeneracies of the modes. For example, in (a linear molecule), the three normal modes consist of one symmetric stretch (inactive in but Raman-active), one asymmetric stretch, and one degenerate bending mode. Real molecular vibrations deviate slightly from the ideal harmonic model due to , arising from the asymmetric shape of the actual , which allows for higher-order terms beyond the quadratic approximation. This leads to effects such as , where transitions to higher vibrational states (\Delta v > 1) occur at frequencies lower than integer multiples of the fundamental, and enables weak interactions between modes that would otherwise be forbidden. While these deviations are small for low-energy vibrations, they become more pronounced in excited states and provide subtle corrections to the energy levels relevant for detailed Raman analysis.

Quantum Theory of Inelastic Scattering

The quantum mechanical description of Raman scattering emerges from time-dependent applied to the between and , treating the process as a second-order transition where an incident excites the system to a before a scattered is emitted with a shift corresponding to a vibrational or difference. In this framework, the scattering is because the final state of the differs from the initial state by a small energy quantum, typically on the order of vibrational energies (\hbar \omega_{vib} ~ 100-4000 cm⁻¹), while the is a non-resonant that does not satisfy for a real transition, ensuring the process remains and coherent. This arises from the , where the of the incident couples the state to an excited manifold, followed by a second coupling to the scattered field, leading to no net change in state but a shift in . The is rigorously captured by the Kramers-Heisenberg-Dirac (KHD) dispersion formula, derived using the correspondence principle and later formalized in , which expresses the induced tensor α as \alpha_{fi} \propto \sum_e \left[ \frac{\langle f | \hat{\mu} | e \rangle \langle e | \hat{\mu} | i \rangle}{E_e - E_i - \hbar \omega + i \Gamma} + \frac{\langle f | \hat{\mu} | e \rangle \langle e | \hat{\mu} | i \rangle}{E_e - E_f + \hbar \omega + i \Gamma} \right], where |i⟩ and |f⟩ are the initial and final molecular states (typically vibrational levels in the ground state), |e⟩ sums over excited states, \hat{\mu} is the electric operator, ω is the incident , E denotes energies, and Γ accounts for the finite lifetime of the . The first term in the sum dominates for non-resonant , reflecting absorption to the followed by , while the complex denominator ensures unitarity and causality, with the imaginary part preventing divergences. This formula unifies (elastic) and Raman (inelastic) under a single quantum mechanical umbrella, with the energy shift ΔE = E_f - E_i determining the scattered ω_s = ω - ΔE/ℏ. In Raman spectra, two primary inelastic processes occur: Stokes scattering, where the molecule gains vibrational energy (ΔE = +ℏ ω_{vib}, ω_s < ω) from the photon, populating an excited vibrational state |v'⟩ from initial |v⟩, and anti-Stokes scattering, where the molecule loses energy (ΔE = -ℏ ω_{vib}, ω_s > ω), requiring thermal population of |v'⟩ beforehand. The intensity ratio of anti-Stokes to Stokes lines follows the , I_{anti-Stokes}/I_{Stokes} = (N_{v'}/N_v) = e^{-\hbar \omega_{vib}/kT}, where at (kT ~ 200 cm⁻¹), this factor suppresses anti-Stokes signals for modes above ~500 cm⁻¹ since higher vibrational levels are sparsely populated. This thermal dependence allows to probe temperature via the Stokes-anti-Stokes ratio without assuming beyond the quantum amplitudes. The differential scattering cross-section, quantifying the probability per unit solid angle per scatterer, is given by dσ/dΩ ∝ |α|^2, where |α|^2 averages over the polarizability tensor components, incorporating orientational and tensorial factors for isotropic samples. The scattered intensity is then I_s ∝ I_0 (dσ/dΩ), with I_0 the incident intensity. This expression, derived from the squared KHD amplitude, scales the weak Raman signal (typically 10^{-6} to 10^{-3} of elastic scattering) and highlights the role of molecular polarizability in determining observable intensities. When the incident frequency ω approaches an transition (resonance condition, |E_e - E_i - ℏω| ≲ Γ), the denominator in the KHD formula shrinks, enhancing the cross-section by factors of 10² to 10⁶ through partial of real electronic states, as formalized in Albrecht's vibronic coupling theory where the A-term (Franck-Condon overlap) dominates for allowed transitions. This resonance Raman effect selectively amplifies coupled to the , providing enhanced for studying chromophores while the non-resonant remains for off-resonance cases.

Selection Rules and Symmetry

Polarizability and Transition Rules

Raman scattering occurs when the polarizability of a molecule changes during a vibrational or rotational transition, allowing for inelastic light scattering. Specifically, a vibrational mode is Raman-active if the polarizability tensor \alpha varies with the normal coordinate Q of the vibration, such that \frac{\partial \alpha}{\partial Q} \neq 0. This contrasts with infrared (IR) absorption, where activity requires a change in the dipole moment \mu, i.e., \frac{\partial \mu}{\partial Q} \neq 0. As a result, Raman spectroscopy complements IR by probing modes that do not involve dipole changes, such as symmetric stretches in homonuclear diatomic molecules like N_2 or O_2. In centrosymmetric molecules, the rule dictates that vibrational modes active in Raman scattering are inactive in absorption, and vice versa. This arises from considerations: under inversion through the center of symmetry, the polarizability tensor (a second-rank tensor) remains unchanged, while the (a ) reverses sign. Thus, vibrations that alter the (-active) cannot change the (Raman-inactive) in such molecules, as exemplified by CO_2, where the symmetric stretch is Raman-active but IR-inactive. For pure rotational Raman scattering in linear molecules, the selection rule is \Delta J = 0, \pm 2, where J is the rotational . This leads to spectral branches: the O branch for \Delta J = -2, the Q branch for \Delta J = 0 (absent in linear molecules due to zero intensity), and the S branch for \Delta J = +2. These branches appear as lines spaced by approximately $4B (where B is the rotational constant), providing information on the molecular . Isotopic substitution affects Raman spectra by shifting vibrational frequencies due to changes in , which alter the in the vibrational potential. Heavier isotopes lower the frequency of modes involving those atoms, enabling identification of atomic contributions to specific . For instance, replacing ^{12}C with ^{13}C in molecules shifts C-H stretches, aiding structural elucidation in Raman . The depolarization \rho = \frac{I_\perp}{I_\parallel}, where I_\perp and I_\parallel are the intensities of Raman-scattered perpendicular and to the incident , respectively, reveals mode . Totally symmetric modes yield \rho \approx 0 (polarized scattering), while non-totally symmetric or asymmetric modes give \rho \approx \frac{3}{4} (depolarized), as the of the tensor randomizes the scattered .

Polarization Effects

In Raman scattering, the polarization of the incident and scattered light provides critical insights into the symmetry properties of molecular vibrations. Vibrational modes are classified according to the irreducible representations of the molecule's symmetry, which dictate the allowed components of the Raman tensor. For instance, in molecules belonging to the C_{3v} , such as (NH₃), modes transform as A_1, E, or other representations, determining which tensor elements contribute to the scattering intensity for specific configurations. This classification arises from , where the Raman activity requires the derivative to belong to the same as the quadratic forms of the Cartesian coordinates. The Raman tensor, denoted as \alpha_{ij}, represents the change in molecular with respect to and transforms as a second-rank under the point group operations. It can be expressed as: \alpha = \begin{pmatrix} \alpha_{xx} & \alpha_{xy} & \alpha_{xz} \\ \alpha_{yx} & \alpha_{yy} & \alpha_{yz} \\ \alpha_{zx} & \alpha_{zy} & \alpha_{zz} \end{pmatrix} where \alpha_{ij} = \alpha_{ji}, and the tensor components are constrained by ; for example, totally symmetric modes (e.g., A_1 in C_{3v}) often exhibit an isotropic part, \alpha_{iso} = (\alpha_{xx} + \alpha_{yy} + \alpha_{zz})/3, leading to depolarization ratios close to zero for parallel polarizations. In contrast, antisymmetric modes may have vanishing isotropic components, resulting in higher . These tensor symmetries enable the assignment of vibrational modes by measuring intensities under varied geometries, such as parallel (VV) or perpendicular (VH) configurations. In anisotropic samples, the orientation of the molecule relative to the light polarization significantly affects the observed Raman spectra. Single crystals exhibit mode splitting or intensity variations due to the crystal's , where the Raman tensor is projected onto the laboratory frame, revealing site-specific contributions; for example, in zeolites with MFI , polarized Raman reveals local anisotropies in silicalite-1 crystals. Powdered samples, however, average over all orientations, yielding isotropic spectra that obscure individual tensor components and mimic solution-phase behavior. Circular polarization in Raman scattering is particularly useful for probing chiral molecules, where Raman Optical Activity (ROA) measures the difference in scattering intensity between left- and right-circularly polarized light. ROA arises from the interference between the polarizability and optical activity tensors, providing a sensitive probe of molecular chirality without requiring magnetic fields. For chiral biomolecules, such as proteins, ROA spectra reveal secondary structure details through differential backscattering intensities. Adsorption on surfaces can alter the effective symmetry of molecules, relaxing selection rules and enabling Raman activity for previously inactive modes, which serves as a prerequisite for enhanced scattering processes like SERS.

Experimental Methods

Instrumentation Components

The evolution of light sources in has been pivotal since the introduction of lasers in the , replacing earlier low-intensity mercury arc lamps that required exposure times of several hours for detectable signals. Prior to lasers, mercury lamps provided broadband illumination but suffered from weak monochromatic lines and high background noise, limiting sensitivity. Continuous-wave (CW) and pulsed lasers now dominate, offering high intensity and narrow linewidths essential for exciting weak Raman scattering events. Common laser wavelengths include 532 nm from frequency-doubled Nd:YAG systems, which provides strong Raman signals due to the inverse fourth-power dependence on wavelength but can induce fluorescence in many samples. In contrast, 785 nm diode lasers are widely adopted for their ability to minimize fluorescence interference, particularly in biological and organic materials, while maintaining adequate signal strength. Other options, such as 1064 nm Nd:YAG for Fourier transform (FT)-Raman to further suppress fluorescence, enable analysis of highly fluorescing samples. Sample illumination in modern Raman setups employs focused to maximize interaction volume and signal collection efficiency. objectives, often with high numerical apertures (e.g., 0.9 ), are standard in micro-Raman configurations to achieve diffraction-limited down to 1 μm, allowing localized probing of heterogeneous samples. For remote or measurements, fiber-optic probes deliver light and collect scattered signals, enabling non-contact analysis in hazardous or inaccessible environments. Collection optics are designed to efficiently gather the faint inelastically scattered light while rejecting the intense elastically scattered (Rayleigh) component. Notch filters, with optical densities exceeding 10^5 at the laser wavelength, are positioned post-sample to block in backscattering geometries, where the collection angle is 180° to the excitation direction for optimal signal throughput. Lenses and mirrors direct the light into the spectrometer, with 90° geometries used in some gas-phase or flow-cell setups to reduce background from the excitation source. Detectors have advanced from single-channel photomultiplier tubes (PMTs), which scanned spectra sequentially in early laser-era systems, to multichannel charge-coupled devices (CCDs) that capture the entire spectrum simultaneously, improving speed and signal-to-noise ratios by factors of 10-100. In FT-Raman systems using near-infrared excitation, specialized indium gallium arsenide (InGaAs) detectors handle the longer wavelengths, enabling high-resolution measurements with reduced fluorescence. Post-2000 developments have emphasized portability and enhanced sensitivity through integrated systems. Handheld Raman units, incorporating compact lasers, fiber-optic probes, and miniaturized spectrometers, facilitate field-deployable analysis for and . Tip-enhanced Raman spectroscopy (TERS) integrates Raman with (AFM), using plasmonic tips to achieve nanoscale (10-20 nm) resolution and signal enhancements up to 10^6, as demonstrated in seminal works on .

Detection and Analysis Techniques

Detection and analysis techniques in involve acquiring high-quality spectra and applying processing methods to extract meaningful information from the signals. is a critical parameter, typically achieved in dispersive systems using spectrometers, where resolutions of 1-10 cm⁻¹ are common depending on the groove density (e.g., 1200-1800 grooves/mm) and entrance slit width. These systems disperse the scattered light across a detector, allowing separation of closely spaced vibrational modes, though higher resolutions (below 1 cm⁻¹) require narrower slits that reduce signal intensity. In contrast, (FT)- employs via a , enabling broader spectral coverage (often 50-4000 cm⁻¹) without the limitations of order overlap, while maintaining resolutions as fine as 0.5-4 cm⁻¹ through and zero-filling of the interferogram. Noise sources pose significant challenges to spectral fidelity, with fluorescence background from the sample often overwhelming the weak Raman signal, particularly when using visible excitation lasers. This broad, slowly decaying emission arises from electronic transitions and can be mitigated by shifting to near-infrared (NIR) lasers (e.g., 785 nm or 1064 nm), which excite below typical absorption bands in biological or organic materials, reducing fluorescence by orders of magnitude while preserving Raman scattering efficiency. Cosmic rays, high-energy particles causing sharp spikes in charge-coupled device (CCD) detectors, represent another intermittent noise source and are subtracted post-acquisition using algorithms that identify and interpolate over these artifacts. Baseline correction techniques, such as polynomial fitting or asymmetric least squares, further remove sloping fluorescence backgrounds by estimating and subtracting a smooth continuum, ensuring accurate peak intensities and positions. Multivariate analysis enhances interpretation of complex Raman spectra, where reduces dimensionality by identifying variance-dominant components, facilitating peak in overlapping regions through loading plots that highlight contributing spectral features. complements this by modeling individual peaks with or Voigt profiles to resolve blended modes and assign them to specific vibrational or rotational transitions, often iteratively refining parameters like peak width and position for quantitative mode assignment. ensures accuracy, commonly using the sharp silicon peak at 520 cm⁻¹ as a standard for dispersive systems, or neon emission lines from a calibration lamp for absolute scale verification across the spectral range. For dynamic processes, time-resolved Raman spectroscopy captures transient spectra on femtosecond to picosecond timescales using streak cameras, which sweep images temporally to achieve resolutions below 1 ps, or gated detectors like intensified CCDs with microchannel plates for sub-nanosecond gating. These techniques synchronize detection with ultrashort pump-probe pulses, enabling observation of vibrational coherences or reaction intermediates while rejecting longer-lived .

Nonlinear and Coherent Variants

Stimulated Raman Scattering

Stimulated Raman scattering () is a nonlinear optical process in which an intense pump interacts with a medium to coherently amplify a weaker Stokes beam at a lower frequency, corresponding to a vibrational or rotational transition of the molecules. This amplification arises from the parametric coupling between the pump, Stokes, and material excitations, where the pump photons are converted into Stokes photons and phonons, leading to exponential growth of the Stokes intensity. The process requires phase matching to ensure efficient momentum conservation, typically achieved in collinear geometries where the wave vectors of the pump and Stokes beams align closely with the phonon wave vector. Unlike spontaneous Raman scattering, SRS involves a coherent buildup of the Stokes field, driven by the third-order nonlinear susceptibility \chi^{(3)} of the medium. The Raman gain coefficient g, which quantifies the amplification per unit length, is proportional to the imaginary part of \chi^{(3)} and the pump intensity I_p, expressed as g \propto \operatorname{Im}(\chi^{(3)}) I_p. This gain enables the Stokes field to grow exponentially as I_s(z) = I_s(0) \exp(g I_p z), where z is the propagation distance. For significant amplification to occur, the pump intensity must exceed a threshold condition where g I_p L / 2 > 1, with L being the interaction length in the medium; below this threshold, the process remains weak and dominated by spontaneous scattering. This threshold ensures that the coherent parametric amplification overcomes losses and noise, leading to efficient energy transfer from the pump to the Stokes beam. Efficient demands precise spatial and temporal overlap between the and Stokes beams to maximize the interaction volume and synchronize the photon fields with the molecular response. Spatial overlap requires collinear or tightly focused beams to maintain matching over the interaction length, while temporal overlap ensures that the and Stokes pulses coincide within the time of the vibrational , typically on the order of picoseconds. These requirements are critical in applications such as Raman fiber lasers, where a at 1064 in silica fiber is shifted to a Stokes wavelength of 1117 via the ~13 THz Raman shift, enabling -tunable sources for and sensing. In high-intensity regimes, can induce () due to the intensity-dependent , which broadens the and facilitates . This nonlinear effect, combined with the Raman gain, can lead to the formation of soliton-like structures that propagate without , enabling the generation of ultrashort s for advanced systems. Such in has been demonstrated in fibers and gases, providing a pathway for high-peak-power applications while maintaining beam quality. The inverse Raman effect, also known as inverse Raman scattering, is a nonlinear optical where a strong pump beam at frequency \omega_p interacts with a beam (often referred to as the Stokes beam) spanning a range of frequencies \omega_s, leading to induced in the probe beam at frequencies corresponding to vibrational Raman resonances of the sample. This absorption manifests as a depletion or reduction in the probe intensity at those specific frequencies, allowing the Raman spectrum to be obtained by measuring changes in the probe transmission through the sample. First observed in using a pump and a in liquids like , the effect arises from the stimulated Raman in reverse, where the intense pump field induces stimulated absorption of the probe beam at frequencies corresponding to Raman resonances. Unlike spontaneous Raman scattering, this coherent enables quantitative for challenging samples, such as fluorescing compounds or low-pressure gases, by providing higher sensitivity and reduced interference from background . A prominent related process is coherent anti-Stokes Raman scattering (CARS), a four-wave mixing technique that generates a coherent, blueshifted anti-Stokes signal at frequency \omega_{AS} = 2\omega_p - \omega_s, where the pump and probe beams are typically at the same \omega_p and the Stokes beam is detuned to \omega_s = \omega_p - \Omega, with \Omega matching the vibrational of the sample. The efficiency of CARS relies on phase-matching condition, \mathbf{k}_{AS} = 2\mathbf{k}_p - \mathbf{k}_s, where \mathbf{k} denotes the wave vectors, ensuring constructive and directional emission of the anti-Stokes . First demonstrated in in gases and liquids, CARS produces a laser-like signal that is inherently coherent and polarized, distinguishing it from the inverse Raman effect by generating new rather than depleting existing light. This amplifies the Raman signal through stimulated , offering detection limits far superior to spontaneous methods. Supercontinuum generation represents another coherent Raman-related process, where intense pump pulses in optical fibers or photonic structures initiate broadband cascades of Stokes and anti-Stokes sidebands, extending the spectrum over octaves via Raman gain and modulation instability. In silica photonic crystal fibers, for instance, the process begins with modulation instability seeding noise-like perturbations that are amplified by stimulated Raman scattering, leading to soliton dynamics and dispersive wave generation for ultra-broadband output from visible to near-infrared wavelengths. This phenomenon, extensively studied since the late , enables compact sources for by producing a white-light rich in Raman features. An analogy to these processes is found in stimulated Raman adiabatic passage (STIRAP), a coherent control technique adapted for molecular population transfer using overlapping pump and Stokes pulses with appropriate time delays and frequency chirps to follow the dark state of a three-level \Lambda-system without populating the intermediate excited state. In Raman contexts, chirped pulses facilitate efficient vibrational state transfer in molecules, mirroring the coherent driving in inverse Raman and CARS but emphasizing adiabatic evolution for quantum state manipulation. Compared to spontaneous Raman scattering, these inverse and coherent variants provide signal enhancements of $10^4 to $10^6 times due to stimulated amplification, with offering additional benefits like coherent, directional signal generation; however, it includes a non-resonant background that can be suppressed using techniques such as polarization-sensitive detection in multiplexed setups for improved spectral contrast. Recent advances as of 2025 include super-resolution techniques and AI-enhanced processing for and , improving imaging capabilities in biomedical applications.

Applications and Extensions

Analytical Spectroscopy

Raman spectroscopy plays a central role in analytical spectroscopy for non-destructive chemical identification and material characterization, leveraging the inelastic scattering of light to generate molecular vibrational spectra that serve as unique fingerprints. The typical spectral range spans 400 to 4000 cm⁻¹, encompassing both the fingerprint region (below 1800 cm⁻¹) for skeletal vibrations and higher-wavenumber bands for functional group modes, such as the symmetric and asymmetric C-H stretching vibrations near 2900 cm⁻¹ in hydrocarbons. These peak positions and relative intensities allow for the identification of molecular structures without sample preparation, making it ideal for analyzing complex mixtures in pharmaceuticals, polymers, and environmental samples. To overcome the inherently weak Raman signal and achieve trace-level detection, (SERS) employs nanostructured metallic surfaces, particularly silver (Ag) and (Au) nanoparticles, where resonances amplify the electromagnetic field. This plasmonic enhancement typically yields factors of 10⁶ to 10⁸ for ensemble-averaged spectra, enabling ultrasensitive analysis down to single-molecule detection under optimized conditions, such as in hot spots between closely spaced nanoparticles. SERS has revolutionized analytical applications by providing structural information at attomolar concentrations, though reproducibility depends on substrate uniformity and analyte-substrate interactions. Quantitative analysis via relies on establishing curves by correlating peak areas or heights with concentrations, often using internal standards to normalize for variations. However, matrix effects—such as interference, self-absorption, or reabsorption in heterogeneous samples—can distort signal linearity, necessitating multivariate methods like for robust predictions. These challenges are particularly evident in solid or turbid media, where deviations from Beer-Lambert-like behavior limit accuracy to 5-10% relative error in favorable cases. In situ Raman spectroscopy excels in real-time monitoring of dynamic processes under non-ambient conditions, such as high-pressure geochemical reactions probed via diamond anvil cells (DACs), which compress samples to gigapascal levels while allowing vibrational spectra to track phase transitions and bonding changes in minerals or fluids. For instance, DACs combined with laser heating enable studies of mantle-like conditions, revealing speciation in silicate melts or polymerization. Similarly, in monitoring, in situ setups detect the formation and evolution of oxide layers on metals, such as products on mild steel in saline environments, by identifying characteristic peaks like those of (Fe₂O₃) at ~1300 cm⁻¹. A key strength of Raman in analytical is its sensitivity to polymorphism, where subtle differences in alter vibrational modes, enabling distinction between allotropes like and in carbon materials. exhibits a sharp triply degenerate peak at 1332 cm⁻¹ due to its sp³-hybridized , while shows a prominent G-band at 1580 cm⁻¹ from in-plane sp² vibrations, allowing rapid identification in or without . This capability extends to pharmaceuticals, where polymorphic forms impact , with Raman providing conformational insights governed by change selection rules.

Imaging and Sensing Technologies

Confocal enables label-free, non-destructive three-dimensional chemical mapping of biological samples at sub-micron spatial resolutions, typically achieving diffraction-limited performance around 0.5–1 μm laterally and 1–2 μm axially through confocal pinhole rejection of out-of-focus light. This technique combines with confocal to generate hyperspectral images that reveal molecular compositions, such as distributions in cell membranes or protein accumulations in tissues, without requiring exogenous labels or staining. For instance, in living tissues, it has been applied to visualize metabolic changes in cancer cells, providing volumetric reconstructions that correlate biochemical signatures with cellular . Raman optical activity (ROA) extends standard Raman spectroscopy by measuring differential scattering intensities from chiral molecules under circularly polarized light, offering exquisite sensitivity to three-dimensional molecular structures and enabling chiral discrimination in biomolecules. In protein analysis, ROA spectra distinguish secondary structural elements like α-helices, β-sheets, and random coils through characteristic band patterns in the amide I and III regions, as demonstrated in studies of globular proteins such as hen egg-white lysozyme. For DNA and RNA, ROA reveals backbone conformations and base stacking interactions, facilitating insights into secondary structures in solution without the need for crystallization or isotopic labeling. This technique's development since the 1990s has positioned it as a complementary tool to circular dichroism for biomolecular stereochemistry. Stand-off Raman detection employs lidar-like systems to identify hazardous materials such as explosives and drugs from distances up to several meters, minimizing operator risk in security applications. These systems often utilize near-infrared excitation at 1064 nm from Nd:YAG lasers to suppress sample , which can overwhelm Raman signals in organic compounds like 2,4,6-trinitrotoluene or , thereby enhancing signal-to-noise ratios for trace detection. Multispectral imaging variants capture spatially resolved Raman spectra across wavelengths, allowing pixel-by-pixel matching to reference libraries for single-particle identification of or dinitrotoluene at ranges exceeding 10 m under controlled conditions. Raman biosensing leverages fiber-optic probes for in vivo or point-of-care applications, such as non-invasive glucose monitoring in fluids, where surface-enhanced Raman scattering (SERS) substrates on probe tips amplify weak signals from glucose's C-H and C-O vibrations at concentrations as low as 1–10 mM. These label-free approaches exploit intrinsic molecular fingerprints, avoiding enzymatic intermediaries for direct quantification in diabetic management. Similarly, for pathogen identification, fiber-coupled Raman systems enable rapid, culture-free detection of bacteria like or by analyzing whole-cell spectral profiles, achieving sensitivities down to 10³–10⁵ colony-forming units per milliliter through multivariate analysis of lipid and protein bands. Multimodal integration of with and facilitates correlative that overlays chemical, structural, and functional data for comprehensive biological analysis. In cellular studies, Raman provides label-free metabolic mapping (e.g., distribution), while highlights specific biomarkers like GFP-tagged proteins, and adds ultrastructural details at nanometer resolution, as seen in investigations of remodeling in tissues. This hybrid approach enhances diagnostic accuracy in by correlating Raman-detected with fluorescent markers and SEM-visualized membrane disruptions, without sample sectioning.

Emerging and Specialized Uses

Tip-enhanced Raman spectroscopy (TERS) integrates scanning tunneling microscopy () tips with Raman scattering to achieve nanometer-scale , enabling the study of single-molecule dynamics and chemical structures at the atomic level. By confining electromagnetic fields to the apex of a metallic tip, TERS enhances Raman signals by factors exceeding 10^4, allowing detection of vibrational modes from individual molecules without ensemble averaging. This technique has revealed transient conformational changes and reaction pathways in molecules like brilliant cresyl blue and perylenetetracarboxylic dianhydride (PTCDA), where time-resolved spectra capture sub-second dynamics influenced by substrate interactions. Applications in include probing plasmonic hotspots and molecular junctions, providing insights into charge transfer and bonding at interfaces. Raman thermometry utilizes the intensity ratio of anti-Stokes to Stokes Raman bands to map distributions in semiconductors with sub-micron , offering a non-contact for thermal characterization in . The ratio follows the relation \frac{I_{as}}{I_s} \propto \exp\left(-\frac{h\nu}{kT}\right), where I_{as} and I_s are the anti-Stokes and Stokes intensities, h is Planck's constant, \nu is the vibrational frequency, k is Boltzmann's constant, and T is ; this Boltzmann factor enables precise thermometry insensitive to effects. In and devices, TERS variants have mapped hotspots during operation, revealing thermal gradients as low as 1 /μm in . This approach supports and in integrated circuits by correlating local heating with performance degradation. Quantum Raman processes in optical cavities leverage Raman interactions to generate entangled photon states, advancing for secure communication and sensing. Through Raman-driven quantum-beat lasers or (QED) systems, four-level atomic schemes produce entangled cavity modes via , achieving fidelities above 90% in controlled setups. In cavity-enhanced configurations, spontaneous Raman scattering couples with parametric processes to create polarization-entangled pairs, useful for quantum networks where entanglement transfer occurs via atomic ensembles. These methods extend to ultrafast , where entangled photons probe molecular coherences with enhanced signal-to-noise ratios compared to classical sources. Machine learning integration has advanced Raman applications, particularly in coherent Raman imaging and vibrational spectroscopy. As of 2025, algorithms enable rapid classification of hyperspectral data, , and predictive modeling for applications in diagnostics and nanomaterial . In space biology, supports non-invasive assessment of plant growth and physiological responses in simulated microgravity environments. A 2025 study demonstrated its use in detecting metabolic changes in plants, aiding development of closed-loop life support systems for future missions. In preservation, enables non-invasive identification of pigments and artifacts, preserving historical integrity while revealing artistic techniques. Portable Raman instruments have distinguished natural —sourced from Afghan mines and valued in paintings—from synthetic variants through characteristic and bands at 547 cm⁻¹ and 259 cm⁻¹. Applications include analyzing 14th-century manuscripts, where spectra confirm use alongside and , aiding authentication and degradation studies without sampling. This technique supports conservation by detecting products in organic binders, ensuring targeted . Raman spectroscopy aboard Mars rovers like Perseverance's SuperCam instrument facilitates remote mineralogical analysis in extraterrestrial exploration, identifying hydrated silicates and carbonates in Jezero Crater since 2021. SuperCam's 532 nm laser excites Raman signals from standoff distances up to 7 m, detecting olivine, pyroxene, and serpentine in deltaic sediments, which indicate past aqueous environments. As of November 2025, after over 1680 sols, it has confirmed diverse igneous and alteration minerals, including carbonates at 1095 cm⁻¹, with recent 2024–2025 observations revealing redox-driven mineral and organic associations, supporting astrobiology goals by mapping organic-mineral associations. This capability enhances sample selection for return missions, providing compositional context for Mars' geological history.

References

  1. [1]
    Raman Spectrometer - The University of Akron
    Raman is a light scattering technique, in which molecular vibrations or collective excitations in solids (phonons) scatter incident light from a high intensity ...
  2. [2]
    Raman Scattering - an overview | ScienceDirect Topics
    Raman scattering is the inelastic scattering of light where the scattered photon's frequency differs from the incident photon, often due to vibrational modes.
  3. [3]
    This Month in Physics History | American Physical Society
    In February 1928, C.V. Raman discovered the Raman effect, a new scattering effect, and observed that the scattered light was polarized.Missing: definition | Show results with:definition
  4. [4]
    C.V. Raman The Raman Effect - Landmark
    Sir CV Raman discovered in 1928 that when a beam of coloured light entered a liquid, a fraction of the light scattered by that liquid was of a different color.Raman Measures the Effect of... · The Laser and Raman...
  5. [5]
    C. V. Raman and the Discovery of the Raman Effect - ADS
    In 1928 the Indian physicist C. V. Raman (1888-1970) discovered the effect named after him virtually simultaneously with the Russian physicists G. S. Landsberg ...
  6. [6]
    The Nobel Prize in Physics 1930 - NobelPrize.org
    The Nobel Prize in Physics 1930 was awarded to Sir Chandrasekhara Venkata Raman for his work on the scattering of light and for the discovery of the effect ...
  7. [7]
    A History of Raman Spectroscopy - BCC Research Blog
    Jan 29, 2019 · Raman spectroscopy was discovered in 1928, initially difficult, revived by lasers in the 1960s, and micro-Raman introduced in the 1970s.Missing: definition | Show results with:definition
  8. [8]
    [PDF] Quantum theory of Rayleigh scattering and Raman ... - bingweb
    Raman scattering or the Raman effect, which is the inelastic scattering of a photon was discovered by C. V. Raman and K. S. Krishnan in liquids, and by G. ...
  9. [9]
    Raman Techniques: Fundamentals and Frontiers - PMC
    The Raman effect originates from the inelastic scattering of light, and it can directly probe vibration/rotational-vibration states in molecules and materials.
  10. [10]
  11. [11]
    18.1: Theory of Raman Spectroscopy - Chemistry LibreTexts
    Oct 24, 2022 · With inelastic scattering, a photon is first absorbed by a particle and then emitted with a change in its energy ( Δ ⁢ E ≠ 0 ); this is called ...
  12. [12]
    Introduction to Raman Spectroscopy Techniques- Oxford Instruments
    Stokes radiation occurs at lower energy (longer wavelength) than the Rayleigh radiation, and anti-Stokes radiation has greater energy. The energy increase or ...
  13. [13]
    5: Raman Spectroscopy - Chemistry LibreTexts
    Mar 16, 2023 · From this virtual state it is possible to have a modulated scatter known as Raman scatter. Raman scatter occurs when there is a momentary ...Learning Objectives · Consider the molecular... · What effect would raising the...
  14. [14]
  15. [15]
    Raman Scattering - HyperPhysics
    Raman scattering is inelastic scattering where photons gain or lose energy, shifting frequency due to vibrational or rotational properties of molecules.Missing: definition | Show results with:definition
  16. [16]
    A New Type of Secondary Radiation | Nature
    A New Type of Secondary Radiation. C. V. RAMAN &; K. S. KRISHNAN. Nature volume 121, pages 501–502 ( ...
  17. [17]
    C.V. Raman and the Raman Effect - Optics & Photonics News
    Raman scattering is described quantum mechanically by the exchange of a phonon (quanta of mechanical energy) between a photon and the non-propagating modes of ...C.V. Raman And The Raman... · Raman's Equipment And... · Attribution Of CreditMissing: definition | Show results with:definition<|control11|><|separator|>
  18. [18]
    A New Radiation: C.V. Raman and the Dawn of Quantum ...
    Mar 3, 2025 · The discovery of the Compton effect in 1923 laid the groundwork for understanding the Raman scattering effect by establishing the concept of ...Missing: definition | Show results with:definition
  19. [19]
    Guide to Raman Spectroscopy - Bruker
    (2) Inelastic scattering, or Raman scattering, occurs when the scattered light has a different energy than the light that initially struck the sample. This ...What Is Raman Spectroscopy? · The Raman Effect · Molecular Vibrations In...
  20. [20]
    Normal Modes - Chemistry LibreTexts
    Jan 29, 2023 · ... 3N - 6 vibrational degrees of freedom whereas a linear molecule has 3N -5 degrees of freedom. Total Degree of Freedom, Translational degrees of ...
  21. [21]
    Number of Vibrational Modes in a Molecule - Chemistry LibreTexts
    Jan 29, 2023 · So there are only 2 rotational degrees of freedom for any linear molecule leaving 3N-5 degrees of freedom for vibration. The degrees of ...
  22. [22]
    5.3: The Harmonic Oscillator Approximates Molecular Vibrations
    Mar 9, 2025 · The harmonic oscillator (red curve) is a good approximation for the exact potential energy of a vibration (blue curve).Missing: x0)^ quantized E_v = hν(
  23. [23]
    7.5 The Quantum Harmonic Oscillator - University Physics Volume 3
    Sep 29, 2016 · The total energy E of an oscillator is the sum of its kinetic energy K = m u 2 / 2 K = m u 2 / 2 and the elastic potential energy of the force U ...Missing: x0)^ quantized hν(
  24. [24]
    5.8: The Energy Levels of a Rigid Rotor - Chemistry LibreTexts
    Oct 5, 2025 · The energy is \(\dfrac {6\hbar ^2}{2I}\), and there are five states with this energy corresponding to \(m_J = +2, \,+1,\, 0,\, ‑1,\, ‑2\). The ...Missing: E_J = BJ(
  25. [25]
    2.8: Rotations of Molecules - Chemistry LibreTexts
    Apr 8, 2021 · \[E_J = \frac{\hbar^2 J(J+1)}{2\mu R^2} = B J(J+1)\]. and are independent of \(M\). Thus each energy level is labeled by \(J\) and is \(2J+1 ...
  26. [26]
    13.9: Normal Modes in Polyatomic Molecules - Chemistry LibreTexts
    Jun 30, 2023 · The normal modes of vibration are: asymmetric, symmetric, wagging, twisting, scissoring, and rocking for polyatomic molecules. Symmetricical ...
  27. [27]
    13.5: Vibrational Overtones - Chemistry LibreTexts
    Jun 30, 2023 · The anharmonic oscillator calculations show that the overtones are usually less than a multiple of the fundamental frequency. Overtones are ...
  28. [28]
    The determination of vibrational anharmonicity in molecules from ...
    This review article considers the origin of vibrational anharmonicity in molecules, and the effects that vibrational resonances have on the anharmonicity ...
  29. [29]
    The quantum theory of the emission and absorption of radiation
    P. A. M. Dirac, Proceedings A, 1927. On the theory of quantum mechanics. Paul Adrien Maurice Dirac, Proceedings A, 1926. The fundamental equations of quantum ...
  30. [30]
    Investigations of the Stokes and anti‐Stokes resonance Raman ...
    May 15, 1984 · A general expression for the Boltzmann factor is also derived in terms of the resonant Stokes and anti‐Stokes scattering intensities. This ...
  31. [31]
    On the Theory of Raman Intensities | The Journal of Chemical Physics
    Search Site. Citation. Andreas C. Albrecht; On the Theory of Raman Intensities. J. Chem. Phys. 1 May 1961; 34 (5): 1476–1484. https://doi.org/10.1063 ...
  32. [32]
    [PDF] Raman Spectroscopy: Theory
    The first observations of the phenomenon were made in liquids by Raman and Krishnan,2 then in crystals by Landsberg and Mandelstam,3 both in 1928. The first ...Missing: original | Show results with:original
  33. [33]
    Electronic structure, Raman tensors, and resonance phenomena in ...
    Mar 1, 2010 · Raman tensors, in particular, have specific symmetries that are related to the symmetries of the vibrations by group theory.4 These symmetries ...
  34. [34]
    Polarized Raman spectroscopy and ab initio phonon calculations
    Jan 3, 2019 · Raman modes of a solid and the Raman tensor associated with it are dependent on the symmetry of the crystal. The form of the Raman tensor will ...Missing: polarizability | Show results with:polarizability<|control11|><|separator|>
  35. [35]
    Local anisotropy in single crystals of zeotypes with the MFI ...
    Dec 16, 2019 · In the present work, we demonstrate that polarised Raman spectroscopy is sensitive also to the anisotropies of single crystals of zeolites and ...<|control11|><|separator|>
  36. [36]
    Raman Optical Activity: A Tool for Protein Structure Analysis
    On account of its sensitivity to chirality, Raman optical activity (ROA), measured here as the intensity of a small, circularly polarized component in the ...
  37. [37]
    Overview of Popular Techniques of Raman Spectroscopy and Their ...
    Mar 11, 2021 · Raman spectroscopy is an outstanding material identification technique. It provides spatial information of vibrations from complex biological samples.Missing: challenges | Show results with:challenges
  38. [38]
    [PDF] Laser Raman spectroscopy as a technique for identification of ...
    However, Raman scattering intensity is inversely proportional to λ4, so 532 nm produces a stronger Raman signal than 785 or 1064 nm. Green lasers (532 nm) are ...
  39. [39]
    A Comprehensive Review of Raman Spectroscopy in Biological ...
    We explore the broad range of Raman applications including molecular structure, biomolecular composition, disease detection, and medication discovery. The Raman ...
  40. [40]
    Laser Wavelength Dependence of Background Fluorescence in ...
    However, since 785 nm has become a standard wavelength in Raman analysis, optical components are more readily available, making 785 nm more attractive as a ...Material And Methods · Fluorescence And Absorption... · Raman Spectroscopy
  41. [41]
    The Effect of Microscope Objectives on the Raman Spectra of Crystals
    Sep 1, 2017 · This article explains the underlying physics of changes in relative intensity and even peak position of certain Raman bands depending on the microscope ...
  42. [42]
    Versatile, efficient Raman sampling with fiber optics
    Artifacts and Anomalies in Raman Spectroscopy: A Review on Origins and Correction Procedures. ... Fiber-based stray light suppression in spectroscopy using ...
  43. [43]
    (PDF) Raman Spectrometry with Fiber-Optic Sampling - ResearchGate
    Aug 6, 2025 · The developed probe uses a single lens for simultaneous illumination and collection of the Raman signal backscattered from the sample surface.
  44. [44]
    A high‐throughput Raman notch filter set - AIP Publishing
    Dec 1, 1990 · The filter set can be used to replace the first two dispersion stages in triple‐stage Raman monochromators commonly employed in multichannel ...
  45. [45]
    Collection optics for a Raman spectrometer based on the 90 ...
    To obtain the Raman spectra of gas media, the spectrometer geometry is most preferable in which scattered radiation is collected at an angle of 90° to the ...
  46. [46]
    Raman scattering collection geometry - Physics Stack Exchange
    Mar 29, 2022 · Most Raman spectroscopy measurements are carried out with the collection optics being at either 180 ∘ (back-scattering) or 90 ∘ from the excitation source.
  47. [47]
    Artifacts and Anomalies in Raman Spectroscopy: A Review on ...
    Oct 8, 2024 · The typical layout of Raman instrumentation, which consists of the following key components: laser source, optics for sample illumination ...
  48. [48]
    From Research to Diagnostic Application of Raman Spectroscopy in ...
    ... photomultiplier tube behind detecting RS, modern set ups usually use a CCD detector. This multichannel way of photon detection (a multichannel array chip ...
  49. [49]
    Portable and Handheld Raman Instruments Open a Multitude of ...
    Feb 26, 2025 · Due to the mobility and ruggedness of the handheld hardware, Raman spectroscopy can be used for police, firefighters, and military applications ...Missing: AFM post- 2000<|control11|><|separator|>
  50. [50]
    Recent Developments in Handheld and Portable Optosensing—A ...
    Aug 6, 2025 · Abstract. Recent developments in portable and handheld opto-chemical analytical instrumentation over the last decade (2000-2010) are reviewed.
  51. [51]
    Toward a New Era of SERS and TERS at the Nanometer Scale
    Feb 6, 2023 · This review are to report new directions and perspectives of SERS and TERS, mainly from the viewpoint of combining their mechanism and application studies.
  52. [52]
    Recent advances in tip-enhanced Raman spectroscopy probe designs
    Dec 27, 2022 · We review the recent advances in tip-enhanced Raman spectroscopy, focusing on the novel probe designs for high signal-to-noise ratios enabled by ...Missing: portable handheld
  53. [53]
    Spectral Resolution and Dispersion in Raman Spectroscopy
    Sep 1, 2020 · A Raman spectrometer's spectral resolution is determined by its spectral dispersion in conjunction with the entrance slit width.Missing: evolution | Show results with:evolution
  54. [54]
  55. [55]
    FT-Raman - an overview | ScienceDirect Topics
    FT-Raman also benefits from advantages inherent to interferometry: high collection efficiency, excellent wavelength precision, easily variable resolution, ...
  56. [56]
    Standard Guide for Testing the Resolution of a Raman Spectrometer
    Dec 21, 2022 · Therefore, data obtained of a narrow Raman band on a FT-Raman system can be used to determine the resolution of a dispersive Raman system. A ...
  57. [57]
    How to Reduce Fluorescence in Raman Spectroscopy - Edinburgh ...
    Dec 6, 2022 · One method to suppress fluorescence during measurements is to select a laser with an excitation wavelength that avoids the spectral window in ...Missing: mitigating | Show results with:mitigating
  58. [58]
    Comparison of principal component analysis and biochemical ...
    PCA was first performed on measured Raman spectra, using princomp function in MATLAB. PCA is a statistical analysis method which can reduce the dimension of the ...
  59. [59]
    [PDF] Curve Fitting in Raman and IR Spectroscopy - Thermo Fisher Scientific
    The goal of spectral curve fitting is to mathematically create individual peaks from a spectrum that, when added together, match the original data.
  60. [60]
    Raman spectroscopy with a 1064-nm wavelength laser as a ...
    Nov 3, 2018 · The calibration of the spectrometer followed three steps: the calibration with known lines of neon lamp, with silicon peak at 520cm−1 520 cm − 1 ...
  61. [61]
    Time-gated Raman spectroscopy – a review - IOPscience
    Oct 27, 2020 · Time-gated Raman spectroscopy (TG RS) uses precise synchronization to collect Raman signals during short laser pulses, rejecting fluorescence. ...
  62. [62]
    Fast Gating for Raman Spectroscopy - PMC - NIH
    These very fast detectors are able to detect single photons and allow gating times as short as 100 ps and less.
  63. [63]
    [PDF] Stimulated Raman Scattering: From Bulk to Nano
    Dec 14, 2016 · Because Im χ(3) is proportional to the Raman cross section, the. SRS signal exhibits the same spectral lineshapes as probed in spontaneous ...
  64. [64]
    Stimulated Raman Scattering
    The gain of stimulated Raman scattering (SRS) is based on the fact that the photon degeneracy of the pump beam ( n ¯ p ) and that of the scattering beam ( n ¯ s ) ...
  65. [65]
    Spectrally Focused Stimulated Raman Scattering (sf-SRS ... - NCBI
    Feb 21, 2023 · Stimulated Raman Scattering (SRS) microscopy is a light-based non-linear imaging method for visualizing a molecule based on its chemical properties.
  66. [66]
  67. [67]
    Numerical study on combined stimulated Raman scattering and self ...
    We numerically examine the dynamics of stimulated Raman scattering combined with self-phase modulation for intense 100-psec-long input pulses in a ...
  68. [68]
    [PDF] Raman Bands - HORIBA
    Raman bands are energy changes (wavelengths) seen when a laser photon scatters by a molecule, creating a unique spectral fingerprint. Examples include bands ...Missing: positions | Show results with:positions
  69. [69]
    Molecular Fingerprint Detection Using Raman and Infrared ...
    In this review article, recent research and development activities for molecular fingerprint detection using Raman and infrared spectroscopy are discussed
  70. [70]
    A Review on Surface-Enhanced Raman Scattering - PMC - NIH
    Surface-enhanced Raman scattering (SERS) has become a powerful tool in chemical, material and life sciences, owing to its intrinsic features.<|separator|>
  71. [71]
    Single-Molecule Surface-Enhanced Raman Spectroscopy - MDPI
    Jun 29, 2022 · In this review, we first elucidate the key concepts of SM-SERS, including enhancement factor (EF), spectral fluctuation, and experimental ...
  72. [72]
    Quantitative Analysis Using Raman Spectrometry - Sage Journals
    Calibration Curve Nonlinearity. A plot of analyte concentration vs. analyte Raman band area should be linear, but significant deviations from linearity can ...
  73. [73]
    Estimating the Analytical Performance of Raman Spectroscopy ... - NIH
    For each concentration, the area under the peak was calculated and used for the calibration curve [38]. Each day, calibration curves were done in triplicate.
  74. [74]
    [PDF] In situ high pressure-temperature Raman spectroscopy technique
    This in situ Raman spec- troscopic technique with laser-heated diamond anvil cells represents an important new method to characterize the structures, chemical ...Missing: corrosion | Show results with:corrosion
  75. [75]
    A novel in-situ Raman spectroscopic cell for aqueous geochemistry ...
    Jun 7, 2023 · The cell is well-suited for moderate pressure applications. Typically, diamond anvil cells can be used for conducting in-situ Raman spectroscopy ...
  76. [76]
    In situ Raman Spectroscopy Monitors the Corrosion of Mild Steel in ...
    Jun 1, 2022 · In this study, in situ Raman spectroscopy was used to detect the formation, growth, and evolution of corrosion inside a salt fog chamber.
  77. [77]
    Graphene Nanocomposites Studied by Raman Spectroscopy
    The Raman spectrum of graphite shows different features compared to diamond. Two distinguishable peaks are revealed at ~1350 cm−1 (D band) and ~1580 cm−1 (G ...
  78. [78]
    Quenchable Transparent Phase of Carbon | Chemistry of Materials
    Broad Raman peaks at 1581 cm-1 and 1355 cm-1 indicative of the presence of sp2 carbon are present. (B) At 240 K the 1581 cm-1 line has become narrower and ...
  79. [79]
    Label-free 3D molecular imaging of living tissues using Raman ...
    Sep 9, 2024 · Confocal Raman microspectroscopy has long been used to provide 3D images of the chemical composition with diffraction-limited resolution and ...
  80. [80]
    Chemical Mapping by Confocal Raman Microscopy - PMC - NIH
    We present an application of confocal Raman microscopy to hyperspectral mapping of nanofibers, where nanoscale features are resolved by means of oversampling.
  81. [81]
    Raman optical activity: a tool for protein structure analysis - PubMed
    This article describes the ROA technique and presents ROA spectra, recorded with a commercial instrument of novel design, of a selection of proteins to ...Missing: discrimination DNA seminal
  82. [82]
    [PDF] The development of biomolecular Raman optical activity spectroscopy
    Among other things, ROA provides information about motif and fold, as well as secondary structure, of proteins; solution structure of carbohydrates; polypeptide ...
  83. [83]
    Resonance-enhanced deep-UV Raman spectroscopy for MDMA ...
    Oct 10, 2025 · ... Raman spectroscopy employing both 785 and 1064 nm laser sources ... UV gated Raman spectroscopy for standoff detection of explosives. Opt ...
  84. [84]
    Stand-off detection of explosives particles by multispectral imaging ...
    A Raman multispectral imaging technique is presented, which can be used for stand-off detection of single explosives particles. A frequency-doubled Nd:YAG ...Missing: drugs | Show results with:drugs
  85. [85]
    Glucose detection through surface-enhanced Raman spectroscopy
    May 8, 2022 · This review article mainly focuses on the development of surface enhanced Raman scattering based glucose sensors in recent 10 years.
  86. [86]
    Optical Methods for Label-Free Detection of Bacteria - PMC - NIH
    Dec 15, 2022 · This paper reviews optical methods such as surface-enhanced Raman scattering spectroscopy, surface plasmon resonance, and dark-field microscopic imaging ...
  87. [87]
    Looking for a perfect match: multimodal combinations of Raman ...
    Aug 12, 2021 · Here, we want to give an overview on the development of multimodal systems that use RS in combination with other optical modalities to improve ...
  88. [88]
    Correlative Raman imaging and scanning electron microscopy for ...
    By stacking and fitting Raman images and SEM images, more information can be obtained as shown in Figure 4A, where the red, blue, green, purple, and yellow ...
  89. [89]
    Tip-enhanced Raman spectroscopy – from early developments to ...
    Jun 22, 2017 · Tip-enhanced Raman scattering (TERS) has been demonstrated to be ideally suited to, eg, elucidating chemical reaction mechanisms, determining the distribution ...
  90. [90]
    Single-Molecule Imaging Using Atomistic Near-Field Tip-Enhanced ...
    May 2, 2017 · This work provides insights into single-molecule imaging based on TERS and Raman scattering of molecules in nanojunctions with atomic dimensions.
  91. [91]
    Tip-enhanced Raman scattering for atomic-scale spectroscopy ... - NIH
    Tip-enhanced Raman spectroscopy (TERS), can be used to probe individual atoms and molecules with atomically confined light.
  92. [92]
    [PDF] Kuball, MHH, & Pomeroy, JW (2016). A Review of Raman
    The Stokes/anti-Stokes temperature measurement method has the advantage that it is rather insensitive to mechanical strain. As this technique is time consuming, ...
  93. [93]
    A Review of Raman Thermography For Electronic and Opto ...
    We review the Raman thermography technique, which has been developed to determine the temperature in and around the active area of semiconductor devices.
  94. [94]
    Two-Dimensional Materials for Raman Thermometry on Power ...
    Sep 1, 2025 · Raman thermometry is a powerful technique for sub-microscale thermal measurements on semiconductor-based devices, provided that the active ...
  95. [95]
    Entangled radiation via a Raman-driven quantum-beat laser
    Sep 10, 2009 · We propose a scheme for the entanglement generation of two cavity modes using a four-level Raman-driven quantum-beat laser (QBL).Missing: scattering | Show results with:scattering
  96. [96]
    Entangled states of cavity fields via the Raman interaction
    A scheme to generate the two-mode field entangled states is proposed, based on the Raman interaction of an effective two-level atom with the pump and Stokes ...
  97. [97]
    Entangled photons enabled ultrafast stimulated Raman ... - Nature
    Jul 15, 2024 · A quantum ultrafast Raman spectroscopy is developed for condensed-phase molecules, to monitor the exciton populations and coherences.
  98. [98]
    Characterization of Lapis Lazuli Pigments Using a Multitechnique ...
    Many of the Raman spectra obtained from areas painted with ultramarine pigments in illuminated manuscript leaves from the 14th century Italian manuscript ...
  99. [99]
    Review of in-situ non- and micro-destructive techniques for pigment ...
    May 28, 2025 · This review summarizes in situ non-destructive and micro-destructive techniques, including spectroscopy, chromatography, microscopy, and phase analysis for ...
  100. [100]
    Applications of Raman Spectroscopy in Art and Archaeology - Rousaki
    Dec 2, 2024 · Raman spectroscopy is considered as one of the most valued and important techniques in the art and archaeology analysis field.
  101. [101]
    SuperCam Raman Spectroscopy Unlocks Jezero Crater's ...
    Mar 31, 2025 · Scientists detail the first 1000 sols of spectroscopic mineral detections on Mars.
  102. [102]
    Perseverance Rover's SuperCam Science Instrument Delivers First ...
    Mar 10, 2021 · This is the first acoustic recording of laser impacts on a rock target on Mars from March 2, 2021, the 12th sol (Martian day) from ...