Fact-checked by Grok 2 weeks ago

Raman spectroscopy

Raman spectroscopy is an analytical technique that utilizes the of monochromatic light to probe the vibrational, rotational, and other low-frequency modes in molecules, providing detailed information about their and composition. This method, based on the Raman effect, enables non-destructive analysis of samples in solid, liquid, or gaseous states, making it invaluable for fields such as , , , and pharmaceuticals. The technique was discovered in 1928 by Indian physicist Sir Chandrasekhara Venkata Raman, who observed that a small portion of scattered light from liquids acquires wavelengths different from the incident light, leading to his award of the in 1930 for this groundbreaking work on light scattering. The fundamental principle of Raman spectroscopy relies on the interaction between incident and molecular vibrations: when a absorbs a , it may enter a and re-emit a of slightly different , resulting in a shift (Stokes or anti-Stokes lines) that corresponds to the difference of the molecular . This occurs because the of the changes during vibration, unlike elastic where the remains unchanged. Typically, a serves as the source, with the scattered light collected and analyzed using a spectrometer to generate a where peaks indicate specific molecular bonds or functional groups. The cross-section for is inherently weak—about 10^{-6} to 10^{-8} times that of —necessitating sensitive detectors and often enhancements like (SERS) for trace detection. Since its inception, Raman spectroscopy has evolved from basic laboratory setups to advanced variants, including confocal for spatial mapping at micrometer resolution and time-resolved techniques for studying ultrafast dynamics. Its non-invasive nature allows applications, such as diagnosing diseases through tissue analysis or monitoring processes in . Key advantages include minimal , high specificity for molecular fingerprints, and compatibility with aqueous environments, which is particularly useful in biological studies where water's Raman signal is less interfering than in . Today, it remains a tool in scientific and industrial quality control, continually advancing with innovations in laser technology and computational analysis.

History

Discovery and Early Experiments

The discovery of the Raman effect emerged in the late amid growing interest in light scattering phenomena, particularly following Arthur Compton's 1923 demonstration of inelastic scattering by electrons, which highlighted the particle-like behavior of light and inspired searches for analogous effects with visible wavelengths. At the time, (IR) spectroscopy was emerging as a tool for probing molecular vibrations but faced limitations, such as strong absorption by and challenges in studying liquids non-destructively, prompting researchers like to explore optical alternatives for vibrational analysis. In early 1928, Indian physicist Chandrasekhara Venkata Raman, working at the Indian Association for the Cultivation of Science in Calcutta, collaborated with his student Kariamanickam Srinivasa Krishnan to investigate scattered from . They focused a beam of sunlight—condensed using a 7-inch —through a filter to isolate monochromatic , such as violet from a cobalt- screen, and directed it into dust-free samples like contained in tubes. The scattered at a 90-degree angle was then examined using a pocket spectroscope and, for precise measurements, a spectrograph to record photographic plates. On February 7, 1928, they observed faint lines shifted from the incident wavelength in the scattered spectrum of , confirming due to molecular interactions; this breakthrough was fully realized by February 28. Shortly thereafter, on February 21, 1928, Soviet physicists Grigory Landsberg and Leonid Mandelstam independently observed the same effect in crystals such as and , though Raman's results were the first to be published. The was announced in a letter to dated February 16, 1928, titled "A New Type of Secondary Radiation," describing the modified frequencies in scattered from various liquids and gases as evidence of a novel molecular distinct from . Raman's work rapidly gained recognition, earning him the in "for his work on the of and for the of the effect named after him." Early experiments faced significant hurdles, including the inherently weak intensity of the Raman-scattered signal—about one in 10^7 photons—which necessitated long photographic exposure times, often hours, under intense illumination to capture discernible spectral lines. Subsequent theoretical interpretations by Hendrik Kramers and , along with , in 1929 provided a quantum mechanical framework for the effect.

Technological Developments

In the decades following the discovery of the Raman effect, early instrumentation relied on mercury arc lamps as excitation sources, which provided broadband illumination but suffered from low intensity and required long exposure times for detectable signals. Photomultiplier tubes (PMTs) emerged as key detectors in the 1950s and 1960s, enabling single-channel detection of weak scattered light, though spectra acquisition often took hours due to the faint Raman signals. A major limitation during this era was interference from , which overwhelmed the much weaker , restricting applications primarily to non-fluorescent samples like and gases. The advent of lasers in the transformed Raman spectroscopy into a practical analytical tool. Continuous-wave argon-ion lasers, introduced around 1964, offered monochromatic, high-intensity excitation at visible wavelengths (e.g., 514.5 nm), dramatically improving signal-to-noise ratios and enabling routine measurements on liquids and solids. The first commercial laser-based Raman spectrometers appeared in , marking the shift from laboratory curiosities to accessible instruments and spurring widespread adoption in chemical analysis. From the 1980s, several innovations addressed fluorescence and sensitivity challenges, enhancing Raman's versatility. Charge-coupled device (CCD) detectors, first applied to Raman in 1987, allowed multichannel detection for faster acquisition times (seconds to minutes) compared to scanning PMTs, revolutionizing spectral throughput. Holographic notch filters, developed in the late 1980s and commercialized around 1990, provided superior rejection of while transmitting Stokes-shifted Raman light, enabling back-illumination geometries without double monochromators. Fourier-transform (FT) Raman spectroscopy, pioneered in 1986 using near-infrared excitation (e.g., 1064 nm from Nd:YAG lasers), minimized in biological and organic samples by leveraging lower energies that avoid electronic transitions. A key milestone was the development of confocal Raman microscopy in the 1980s, which incorporated pinhole apertures for depth-resolved imaging with micrometer spatial resolution, laying the foundation for spatially selective analysis. Post-2000 advancements focused on portability and integration, broadening Raman's field applicability. Handheld spectrometers, miniaturized with compact lasers and solid-state detectors, emerged in the early , achieving battery-powered operation for on-site while maintaining above 10 ⁻¹. Fiber-optic probes enabled in hazardous environments, with flexible bundles delivering and collecting scattered over distances up to several meters. with microscopy systems advanced chemical , combining Raman with confocal setups for three-dimensional mapping of heterogeneous materials. accelerated through companies like HORIBA, which delivered early laser systems in 1968 and launched the first true confocal (LabRAM) in 1993, and Renishaw, which introduced its inaugural Raman products in 1992, driving market growth in research and industry. In the 2020s, (AI) has further refined Raman systems for high-sensitivity, rapid acquisition. algorithms, particularly deep neural networks, perform real-time and baseline correction on raw spectra, enabling faster processing (sub-second) even from low-signal data, thus supporting high-throughput applications without compromising accuracy.

Fundamental Principles

Theory of Raman Scattering

The classical theory of Raman scattering describes the process as arising from the modulation of a molecule's polarizability by its vibrational motion. When monochromatic light interacts with a molecule, it induces an oscillating dipole moment proportional to the molecular polarizability tensor, \alpha. Vibrations alter the electron distribution, causing \alpha to vary with the normal coordinate Q of the vibration, which leads to scattered light with frequency shifts corresponding to the vibrational frequencies. In the quantum mechanical framework, Raman scattering involves virtual transitions to intermediate energy states, where the incident excites the molecule without populating real electronic states, followed by emission of a scattered with altered . This is captured by the Kramers-Heisenberg-Dirac , which expresses the scattering cross-section as a over these virtual states, incorporating the incident and scattered light frequencies along with molecular transition moments. The , derived initially by Kramers and Heisenberg in 1925 and extended by Dirac in 1927, provides the quantum basis for the intensity and polarization of the scattered light. The intensity of Raman scattering for a given vibrational mode is proportional to the square of the derivative of the with respect to the normal coordinate: I \propto \left( \frac{d\alpha}{dQ} \right)^2 This relationship, central to Placzek's , quantifies how vibrational distortions that significantly change \alpha produce stronger Raman signals. Unlike elastic , where the scattered 's energy equals the incident 's, Raman scattering is inelastic, involving energy transfer \Delta E = h \nu_m between the and , where \nu_m is the molecular vibrational , resulting in shifted scattered light. Raman selection rules require a nonzero change in polarizability (d\alpha/dQ \neq 0) for a vibrational transition to be active, contrasting with infrared spectroscopy, which demands a change in the dipole moment. This distinction arises from the second-order nature of Raman as a polarizability-driven process versus the first-order dipole interaction in IR absorption. Fluorescence can compete with as an alternative relaxation pathway following virtual , often overwhelming the weaker Raman signal due to its resonant nature.

Raman Shift and Spectral Features

The Raman shift quantifies the energy difference between the incident and scattered photons in Raman spectroscopy, expressed as the wavenumber difference \Delta \nu = \nu_0 - \nu_s in units of cm⁻¹, where \nu_0 is the of the and \nu_s is that of the scattered . This shift directly corresponds to the energy levels of molecular or rotational transitions, providing a spectroscopic signature independent of the . Raman scattering manifests as Stokes or anti-Stokes shifts relative to the incident frequency. In Stokes scattering, the scattered photon has lower energy (longer wavelength) than the incident photon, as the molecule absorbs energy to transition from its ground vibrational state to an excited state. Anti-Stokes scattering involves higher-energy scattered photons (shorter wavelength), where the molecule starts in an excited vibrational state and relaxes to the ground state, transferring energy to the photon. These processes enable the probing of vibrational modes, with Stokes shifts being more commonly observed due to the prevalence of molecules in the ground state at typical temperatures. The relative intensities of Stokes and anti-Stokes peaks are governed by the thermal population of vibrational states, following the . Specifically, the intensity ratio is given by \frac{I_\mathrm{AS}}{I_\mathrm{S}} \propto e^{-hc\Delta\nu / kT}, where h is Planck's constant, c is the , k is , \Delta\nu is the Raman shift, and T is the absolute temperature; this results in anti-Stokes signals being significantly weaker, especially for higher shifts or lower temperatures. Key spectral features in Raman spectra include the fingerprint region, spanning approximately 400 to 1800 cm⁻¹, where complex patterns of fundamental vibrational bands provide unique identifiers for molecular structures, akin to a chemical "fingerprint." Beyond this, overtones—multiples of fundamental vibrational frequencies—and combination bands, arising from interactions between multiple modes, appear at higher wavenumbers (often above 1800 cm⁻¹), offering additional insights into anharmonic effects and molecular coupling, though they are typically weaker in intensity. Raman shifts are sensitive to environmental conditions, with temperature increases generally causing a red shift (to lower wavenumbers) due to and anharmonic vibrations, while induces a blue shift (to higher wavenumbers) by compressing the lattice and strengthening bonds. Isotopic substitution further modulates shifts through changes in atomic mass, altering the in vibrational frequencies and enabling precise studies of . For example, replacing ^{16}O with ^{18}O in \alpha-Fe_2O_3 shifts Raman modes to lower wavenumbers under . In practical Raman measurements, accurate determination of shifts relies on calibration with standards like the transverse optical phonon mode of , which produces a sharp peak at 520 cm⁻¹ under ambient conditions, serving as a reliable reference for accuracy across instruments.

Instrumentation

Excitation Sources

In Raman spectroscopy, excitation sources primarily consist of lasers, which provide the monochromatic and coherent illumination necessary for efficient detection. These sources must deliver high-intensity light with narrow linewidths to resolve subtle Raman shifts, typically operating in continuous-wave () mode for standard measurements, though pulsed lasers are employed in time-resolved or nonlinear variants like coherent anti-Stokes (). Common laser types include gas lasers, solid-state lasers, and semiconductor diode lasers, each offering distinct advantages in wavelength versatility and operational efficiency. Gas lasers, such as helium-neon (He-Ne) at 632.8 nm and argon-ion (Ar+) at 488 nm or 514.5 nm, were historically prevalent due to their excellent beam quality and stability, but they are now less favored owing to high power consumption, short lifetimes, and maintenance needs. Solid-state lasers, exemplified by frequency-doubled Nd:YAG at 532 nm or fundamental Nd:YAG at 1064 nm, provide robust, high-power output (up to several watts) with good efficiency, making them suitable for applications requiring deep tissue penetration or resonance enhancement, though they can pose risks of sample photodamage at shorter wavelengths. Diode lasers, particularly at 785 nm in the near-infrared (NIR), dominate modern setups for their compactness, reliability, low cost, and long operational life (>10,000 hours), enabling portable instruments while minimizing fluorescence interference from biological samples. Wavelength selection critically influences scattering efficiency, which scales inversely with the of the (∝ 1/λ⁴), favoring shorter wavelengths for stronger signals but risking higher and damage. The 785 nm diode strikes an optimal balance for most routine analyses, offering reduced autofluorescence compared to visible sources like 532 nm (used for surface-enhanced , SERS, on inorganics) while maintaining compatibility with silicon-based detectors; in contrast, 1064 nm Nd:YAG lasers are reserved for highly fluorescent samples, such as organics or biomolecules, despite their weaker signals necessitating Fourier-transform (FT) detection. (UV) excitations around 244-257 nm enable Raman for enhanced sensitivity in specific chromophores but demand low powers (<1 mW) to avoid photodegradation. Power levels typically range from 1 to 100 mW at the sample, sufficient to generate detectable scattering without excessive heating, though higher powers (up to 500 mW) may be used in microscopy for localized enhancement. Stability is paramount, with intensity fluctuations <0.1% and wavelength drift <0.1 cm⁻¹ over hours ensuring reproducible spectra; volume Bragg grating (VBG)-stabilized diode lasers achieve this through spectral narrowing to linewidths <0.01 cm⁻¹. Emerging sources include quantum cascade lasers (QCLs) for tunable mid-infrared Raman, facilitating direct vibrational fingerprinting with high spectral purity, and multi-wavelength systems (e.g., dual 785/830 nm diodes) for excitation profile mapping to distinguish fluorescence from Raman signals. Safety considerations are critical due to the high coherence and intensity of these sources, which pose significant eye hazards—particularly invisible NIR wavelengths like 785 nm or 1064 nm that bypass blink reflexes—necessitating Class 3B/4 laser protocols, enclosures, and interlocks; sample burning risks also arise at powers >100 mW on dark or absorbing materials.

Detection and Spectroscopy Systems

In Raman spectroscopy, detection and spectroscopy systems are essential for collecting and analyzing the inelastically scattered light to generate spectra. Dispersive spectrometers, the most common type, employ grating-based monochromators to separate wavelengths of scattered light. These systems use diffraction gratings to disperse the light, with spectral resolution primarily determined by the grating's dispersion (typically measured in nm/mm or cm⁻¹/mm) and the width of the entrance and exit slits. For instance, higher groove densities in gratings (e.g., 1200–2400 lines/mm) and narrower slits (10–100 μm) enable resolutions of 1–10 cm⁻¹, allowing separation of closely spaced Raman bands. The focal length of the spectrometer also influences resolution, with longer focal lengths providing better performance but increasing instrument size. Detectors convert the dispersed photons into electrical signals for spectral readout. (CCD) arrays dominate modern dispersive Raman systems due to their multichannel detection, high (up to 95% in the visible range), and low noise, enabling simultaneous acquisition across the spectrum for improved signal-to-noise ratios. Photomultiplier tubes (PMTs) serve as single-channel detectors in scanning monochromators, offering high sensitivity for weak signals but requiring sequential wavelength scanning, which limits throughput. For (NIR) excitation (e.g., 785–1064 nm), (InGaAs) detectors are preferred, providing sensitivity up to 1700 nm while minimizing interference, though with lower than silicon-based CCDs. Fourier transform (FT)-Raman systems utilize a Michelson interferometer to encode the scattered light into an interferogram, which is then Fourier-transformed to yield the spectrum, offering advantages over dispersive methods for certain applications. The Michelson setup, with a fixed and movable mirror, provides a broader spectral range (often 50–4000 cm⁻¹) and the Fellgett (multiplex) advantage, where all wavelengths contribute to the signal simultaneously, enhancing signal-to-noise by the square root of the number of resolution elements compared to sequential dispersive detection. This is particularly beneficial for low-light samples, though FT-Raman typically pairs with NIR lasers for reduced fluorescence. Resolution and dynamic range in these systems involve inherent trade-offs, as finer resolution (e.g., via slits or gratings) reduces light throughput, potentially lowering sensitivity. CCD detectors mitigate noise through thermoelectric cooling (to -60°C or lower), which suppresses dark current (from ~10⁴ e⁻/pixel/s at room temperature to <1 e⁻/pixel/s), extending integration times and boosting dynamic range to 50,000:1 or more without saturation. In FT-Raman, resolution depends on the maximum mirror travel (e.g., 2 cm for 0.5 cm⁻¹), but the system's higher étendue allows greater light collection. Modern Raman instruments increasingly integrate dispersive and FT capabilities in hybrid configurations for enhanced versatility, allowing users to switch between modes based on sample properties like fluorescence propensity. For example, platforms combining visible-range dispersive CCD detection with 1064 nm FT-Raman interferometry enable both high-resolution structural analysis and fluorescence-free measurements on the same setup. These systems maintain compatibility with various laser wavelengths while optimizing for diverse applications in materials and biological analysis.

Optical Components

Optical components in Raman spectroscopy instrumentation are essential for directing, focusing, and purifying the excitation and scattered light to achieve high signal quality and minimal interference. These components include lenses, mirrors, filters, and specialized optics that manage the beam path while accommodating the weak nature of Raman signals relative to elastic . Beam delivery systems primarily utilize lenses to focus the laser excitation onto the sample and collect the scattered light. In micro-Raman setups, high-numerical-aperture microscope objectives, such as 50× lenses, achieve spot sizes as small as 5 μm, enabling precise illumination of microscopic regions. Fiber optics facilitate remote sensing by transmitting the laser beam and collecting Raman signals over distances exceeding 10 m, which is particularly useful for in-situ applications like forensic analysis. Filters play a critical role in rejecting Rayleigh-scattered light, which is orders of magnitude stronger than Raman signals. Notch or edge filters, often with optical densities greater than 6, effectively block the excitation wavelength and anti-Stokes lines while transmitting Stokes-shifted Raman light. Long-pass edge filters specifically isolate lower-energy Stokes Raman signals by attenuating higher-frequency components, enhancing spectral clarity. Bandpass filters may also be employed to purify the incoming laser beam, reducing noise from unwanted wavelengths. For polarized Raman configurations, polarizers and analyzers are integrated into the optical path to control the polarization state of the incident and collected light. These components, such as thin-film polarizers or polarizing prisms, allow selective transmission based on orientation, supporting studies of molecular symmetry without altering the core scattering geometry. Confocal optics improve spatial resolution by incorporating pinholes that act as spatial filters, rejecting out-of-focus light and confining analysis to the focal plane. A typical pinhole size of 10–50 μm can project to effective resolutions of 100–500 nm when paired with high-magnification objectives (e.g., 100×, NA 0.9), achieving depth resolutions around 630–680 nm full width at half maximum (FWHM). Materials and design features of these components minimize losses and aberrations. Anti-reflective coatings on lenses and mirrors reduce surface reflections to below 0.5%, optimizing light throughput across visible and near-infrared wavelengths used in Raman excitation. Achromatic lenses, composed of paired elements like crown and flint glass, correct chromatic dispersion to maintain focus over a broad spectral range, ensuring consistent performance for both excitation and detection. Glass substrates are commonly used for mirrors and lenses due to their compatibility with Raman wavelengths.

Data Analysis

Spectrum Acquisition

Raman spectroscopy is inherently non-destructive, allowing direct analysis of samples without the need for extensive preparation, which preserves molecular integrity and avoids introducing artifacts from processing. For solids, including amorphous, crystalline, mineral, or organic materials, minimal preparation is required—no polishing, coating, or sectioning is necessary, enabling in situ examination even in transparent samples for depth profiling. Liquids and gases can also be analyzed with little to no preparation; aqueous solutions and gas phases, such as or in inclusions, are suitable for direct measurement, though opaque or colored liquids may require dilution to mitigate heating effects. However, in anisotropic crystals or oriented polymers, sample orientation relative to the laser polarization influences relative peak intensities due to the directional nature of Raman scattering, necessitating consistent alignment for reproducible results. During spectrum acquisition, key parameters are optimized to balance signal intensity, sample safety, and data quality. Exposure time is typically maximized for weak scatterers to enhance the signal-to-noise ratio (SNR), with longer single exposures (e.g., 30 seconds) often preferred over multiple short ones for the same total duration to minimize noise accumulation. Laser power is adjusted finely, starting at low levels (e.g., tenths of milliwatts) to prevent photodegradation or heating, particularly in biological or hydrated samples, while using the full available power (milliwatts range) for robust signals. The number of scans or accumulations is selected based on the sample's Raman cross-section, with multiple acquisitions (e.g., 2–10) averaged to improve SNR and reject transient noise, guided by automated software for efficiency. Common experimental geometries facilitate flexible sample handling and signal collection. In the 90-degree setup, the scattered light is collected perpendicular to the incident laser beam, which is advantageous for avoiding direct laser interference and suitable for gas or liquid cells. Backscattering geometry, where detection occurs in the direction opposite to the excitation, is widely used in microscopic configurations for its simplicity and high collection efficiency, especially with focused beams on solid surfaces. Transmission setups, though less common, involve passing the laser through the sample and collecting on the opposite side, ideal for thin, transparent films or flow cells to maximize path length in dilute samples. Artifacts can compromise spectral fidelity, requiring immediate attention during acquisition. Cosmic rays produce sharp, narrow spikes from high-energy particles hitting the detector, distinguishable from Raman peaks by their asymmetry; removal involves algorithms like wavelet transforms, median filtering, or nearest-neighbor comparison from multiple scans to interpolate affected pixels without distorting underlying signals. Baseline drifts, often from fluorescence or instrumental background, are addressed preliminarily through choice of excitation wavelength or simple polynomial fitting to subtract sloping contributions, ensuring cleaner raw data before further processing. Calibration standards ensure accurate intensity and shift measurements. Internal standards, such as solvent peaks (e.g., water's O-H band at ~3400 cm⁻¹), provide relative referencing within the sample matrix for quantitative corrections. External calibrants, like NIST SRM 2246a (a chromium-doped borosilicate glass), are measured separately under identical conditions to generate correction curves for relative intensity, placed at the sample position with low laser power (≤5.9 mW macro) to validate system response across the spectral range.

Interpretation and Processing

Interpreting Raman spectra involves processing raw data to isolate meaningful signals from noise, fluorescence, and instrumental artifacts, enabling the extraction of chemical composition, molecular structure, and phase information. This processing typically begins with preprocessing steps to enhance spectral quality, followed by analytical techniques to identify and quantify features such as peak positions, intensities, and widths, which correspond to vibrational modes of the sample. For instance, Raman shifts can be assigned to specific vibrational modes like stretching or bending, providing insights into molecular bonds. Accurate interpretation requires careful consideration of the spectrum's resolution and potential errors to avoid misidentification. Baseline subtraction is a fundamental preprocessing step to remove sloping backgrounds caused by fluorescence or instrument response, ensuring that subsequent analysis focuses on true Raman signals. Polynomial fitting methods model the baseline as a low-order polynomial curve fitted to spectral regions devoid of peaks, which is then subtracted from the raw spectrum; this approach is widely used due to its simplicity and effectiveness for moderately curved baselines. The rubberband method, an iterative algorithm, treats the spectrum as a string pulled taut around peaks, estimating the baseline by connecting local minima and is particularly useful for complex, uneven backgrounds without assuming a specific functional form. Both techniques require optimization to prevent over-subtraction, which can distort peak areas. Peak fitting deconvolves overlapping bands in Raman spectra to resolve individual components, often using Lorentzian or Gaussian profiles that reflect homogeneous or inhomogeneous broadening mechanisms, respectively. Lorentzian functions are preferred for Raman peaks arising from lifetime-limited processes, while Gaussian profiles account for Doppler or instrumental broadening; hybrid Voigt profiles combine both for more accurate fits in real samples. Deconvolution algorithms, such as least-squares optimization, iteratively adjust peak parameters (position, width, height) to minimize residuals between the model and data. Quantification of species concentrations relies on band area ratios, normalized to internal standards, as peak areas are proportional to the number of scattering molecules under non-resonant conditions. For complex mixtures where individual peaks overlap extensively, multivariate analysis techniques decompose spectra into latent variables to uncover underlying patterns. Principal component analysis (PCA) reduces dimensionality by identifying principal components that capture the maximum variance, facilitating visualization of spectral similarities and differences in score plots. Partial least squares (PLS) regression extends this for predictive modeling, correlating spectral features with concentration variables in calibration sets, which is essential for quantitative analysis of multi-component systems. These methods handle collinearity and noise effectively but demand robust preprocessing to avoid artifacts propagating into the models. Database matching accelerates identification by comparing processed spectra against reference libraries, leveraging similarity metrics like correlation coefficients or Euclidean distances. The , containing over 5,000 mineral Raman spectra with associated crystallographic data, enables automated matching for geological samples, achieving high accuracy when peak positions align within 2-5 cm⁻¹. Advanced tools like the integrate search algorithms to rank candidates, incorporating orientation and polarization effects for refined results. Emerging artificial intelligence (AI) and machine learning (ML) approaches automate classification and interpretation, addressing challenges in real-time analysis of diverse datasets. Neural networks, particularly convolutional neural networks (CNNs), learn hierarchical features from spectra for tasks like phase identification, outperforming traditional methods in accuracy for noisy or low-signal data. Recent integrations, such as deep learning frameworks trained on large spectral repositories, enable end-to-end processing from raw acquisition to molecular assignment, with applications in portable devices for on-site detection. These models bridge gaps in manual interpretation by handling non-linear relationships but require validation datasets to ensure generalizability. Common error sources in interpretation include overfitting in multivariate models, where excessive components capture noise rather than signal, leading to poor predictive performance on new data; this is mitigated by cross-validation techniques. Resolution limits, dictated by spectrometer slit width and grating dispersion, can broaden peaks below 1 cm⁻¹, obscuring fine structure and introducing uncertainties in linewidth measurements up to 20% for closely spaced bands.

Applications

Chemistry and Materials Science

Raman spectroscopy is widely employed in chemistry and materials science for its ability to provide detailed vibrational information that reveals molecular structures, phase compositions, and mechanical properties without sample destruction. In chemical analysis, it identifies functional groups through characteristic scattering bands, while in materials characterization, it detects subtle changes in lattice vibrations indicative of polymorphism, stress, or conformational order. This technique's specificity arises from the , which corresponds to energy differences between incident and scattered photons, allowing correlation with molecular bonds and crystal symmetries. For molecular identification, Raman spectroscopy excels at detecting functional groups via their unique vibrational signatures. For instance, the C=C stretching mode in alkenes produces a prominent band between 1600 and 1700 cm⁻¹, enabling rapid characterization of unsaturated hydrocarbons in organic compounds. Similarly, C-H stretching vibrations appear around 2800-3000 cm⁻¹, and carbonyl (C=O) groups show strong signals near 1700-1750 cm⁻¹, facilitating the analysis of complex mixtures like pharmaceuticals or reaction products. These assignments are based on established correlation tables that link spectral features to bond types, supporting qualitative and quantitative assessments in synthetic chemistry. In the study of polymorphism and phase transitions, Raman spectroscopy distinguishes between crystal forms by their distinct lattice and molecular vibrations. A classic example is the differentiation of diamond and graphite: diamond exhibits a sharp peak at 1332 cm⁻¹ due to its sp³-hybridized carbon lattice, whereas graphite shows characteristic G and D bands at approximately 1580 cm⁻¹ and 1350 cm⁻¹, respectively, reflecting its layered sp² structure. This sensitivity has been crucial for analyzing carbon polymorphs in geological samples and synthetic materials, where phase transitions under pressure or temperature cause predictable shifts or splitting of these bands. Stress and strain analysis in materials like semiconductors relies on Raman's responsiveness to lattice deformations. In silicon, the unstressed transverse optical phonon mode appears at 520 cm⁻¹; under tensile strain, this peak shifts to lower wavenumbers (red shift) by about 2-3 cm⁻¹ per percent strain, while compressive stress induces a blue shift. This correlation allows non-invasive mapping of residual stresses in microelectronic devices, such as those arising from thermal expansion mismatches in silicon-on-insulator structures, aiding optimization of semiconductor performance. Quantitative models relate the shift magnitude to stress components using tensor analysis, with experimental validation confirming accuracies within 10 MPa. In catalysis studies, in-situ Raman spectroscopy monitors surface reactions and active sites in real time under operational conditions. For example, during CO₂ reduction on copper catalysts, it detects adsorbed intermediates like *CO at around 2000 cm⁻¹, revealing mechanistic pathways and catalyst evolution. This approach has elucidated structure-activity relationships in heterogeneous systems, such as vanadium oxide catalysts where isolated V=O species produce bands at 990 cm⁻¹, shifting with coordination changes during propylene epoxidation. By operating at ambient pressure and temperature, it provides insights into transient species that inform catalyst design for industrial processes. For polymers, Raman spectroscopy assesses conformation and crystallinity through band intensity ratios sensitive to ordered versus disordered chain segments. In polyethylene, the ratio of the 1130 cm⁻¹ band (crystalline trans conformation) to the 1060 cm⁻¹ band (amorphous gauche) quantifies crystallinity levels up to 80%, correlating with mechanical properties like tensile strength. This method tracks phase transitions during processing, such as melting where crystalline bands diminish, and has been applied to polyesters like PET, where the 1725 cm⁻¹ C=O stretch intensity varies with chain packing. Such analyses support quality control in polymer manufacturing by linking spectral metrics to macroscopic behavior.

Biology and Medicine

Raman spectroscopy has emerged as a powerful, label-free tool in biology and medicine, enabling non-destructive analysis of biomolecules and tissues based on their vibrational signatures. In biomedical research, it provides insights into molecular compositions without altering samples, facilitating applications from disease diagnostics to drug development. In tissue analysis, Raman spectroscopy excels at detecting cancer through characteristic spectral ratios, such as the nucleic acid-to-lipid ratio, where elevations in nucleic acid peaks around 780–810 cm⁻¹ relative to lipid peaks at ~1445 cm⁻¹ indicate malignant transformations in cells. For instance, studies on breast and skin tissues have demonstrated high sensitivity in distinguishing tumor margins from healthy tissue, achieving accuracies up to 95% in ex vivo samples by analyzing lipid and protein vibrational modes. This approach leverages the technique's ability to map biochemical heterogeneity, aiding surgical guidance and pathology. For live cell imaging, Raman spectroscopy offers non-invasive probing of intracellular components like proteins and DNA, allowing real-time monitoring of cellular processes without fluorescent labels that could perturb dynamics. Coherent Raman variants, such as , have been used to visualize lipid droplets and protein distributions in living cells with subcellular resolution, revealing metabolic changes during processes like . Micro-Raman configurations enhance this by providing cellular-scale spatial detail in a single-pass scan. In pharmaceuticals, Raman spectroscopy assesses drug-polymer interactions and bioavailability by identifying molecular fingerprints of active ingredients in formulations. It detects polymorphic forms and excipient compatibility, as seen in analyses of tablet coatings where shifts in carbonyl stretches (around 1700 cm⁻¹) signal interactions affecting dissolution rates. This has optimized controlled-release systems, improving therapeutic efficacy. Recent advancements in the 2020s have expanded Raman's clinical utility, including ex vivo diagnostics for brain tumors via fiber-optic probes that classify glioma subtypes with over 90% accuracy based on collagen and nucleic acid signals. Integration with endoscopy enables in vivo applications, such as real-time colorectal cancer screening during procedures. Additionally, AI-enhanced spectral histopathology processes Raman datasets to automate tumor grading, incorporating machine learning models trained on large cohorts for enhanced precision in 2025 clinical trials. Despite these progresses, challenges persist in biomedical Raman applications, particularly water interference that overwhelms weak biomolecular signals and autofluorescence from tissues, which can obscure spectra in the 1000-1800 cm⁻¹ region. Strategies like surface-enhanced Raman scattering mitigate these, but require careful optimization for in vivo reliability.

Art and Cultural Heritage

Raman spectroscopy has become an indispensable tool in the conservation and study of art and cultural heritage due to its non-destructive nature, allowing the identification of materials without damaging irreplaceable artifacts. In particular, it excels at characterizing pigments, binders, and substrates in paintings, sculptures, and manuscripts, providing insights into artistic techniques, historical trade routes, and degradation processes. This technique's ability to probe molecular vibrations enables the detection of both inorganic and organic components, often in situ, which is crucial for museum and archaeological settings where sampling is prohibited. A primary application is pigment identification, where characteristic Raman peaks reveal the chemical composition of colors used by artists across eras. For instance, vermilion (HgS) is readily identified by its strong peak at 253 cm⁻¹, as seen in analyses of ancient manuscripts and Renaissance works. Ultramarine blue, derived from lapis lazuli, shows peaks around 547 and 591 cm⁻¹, helping trace its expensive sourcing from Afghanistan in medieval European illuminations. Other examples include hematite (α-Fe₂O₃) at 225 and 293 cm⁻¹ in red ochres from prehistoric cave art, and emerald green (Cu(C₂H₃O₂)₂·3Cu(AsO₂)₂) in 19th-century murals. These identifications not only document palette evolution but also monitor fading or alteration, such as the darkening of vermilion due to photo-induced conversion to black metacinnabar. In authentication efforts, Raman spectroscopy distinguishes genuine artworks from forgeries by detecting anachronistic materials or mismatched binders. For example, the presence of modern titanium white (TiO₂, peaks at 144, 447, and 612 cm⁻¹), unavailable before the 20th century, has exposed fakes in purportedly ancient Russian avant-garde paintings. Binder analysis, though challenged by fluorescence, can reveal synthetic resins in oil paints, as in cases where 19th-century canvases showed incompatible modern adhesives. This molecular fingerprinting aids provenance verification, such as confirming natural vs. synthetic ultramarine in suspected 18th-century miniatures. Notable case studies illustrate Raman's impact, including the analysis of Vincent van Gogh's paintings, where portable systems identified chrome yellow variants (PbCrO₄) via peaks at ~841 cm⁻¹ (ν₁ CrO₄²⁻) and 400 cm⁻¹ in works like Sunflowers (1888) and Bank of the Seine (1887), revealing his use of sulfate-rich forms prone to darkening. In ancient contexts, Raman examined the 4th-century BC Lakkoma tomb in Macedonia, detecting vermilion and red ochre on a funeral couch, informing Macedonian burial practices. Similarly, a 16th-century Persian manuscript yielded azurite (peaks at 1095 and 1420 cm⁻¹) and malachite, highlighting Qajar dynasty pigment sourcing. Portable Raman instruments facilitate in-situ museum studies, equipped with fiber-optic probes to analyze artifacts without relocation, as demonstrated in the on-site examination of Roman engraved gems using 785 nm excitation to avoid fluorescence. These devices, often combined with low-power lasers (e.g., 1-10 mW), minimize risks to sensitive materials while achieving spatial resolutions down to 1 μm via micro-Raman adaptations. Advantages include no sample preparation and complementarity with for elemental data, enabling layered stratigraphy in cross-sections without destruction—for instance, pairing Raman's molecular specificity with XRF's inorganic mapping in van Gogh's irises. This synergy has revolutionized heritage preservation by providing verifiable, high-resolution insights into cultural objects' composition and history.

Environmental Science

Raman spectroscopy has emerged as a vital tool in environmental science for non-destructive, in situ analysis of pollutants and ecosystems, enabling the identification of chemical signatures in complex matrices like water, air, and soil without extensive sample preparation. Its molecular specificity allows for the detection of trace contaminants at environmentally relevant concentrations, supporting monitoring efforts for sustainability and pollution control. In water quality assessments, Raman spectroscopy excels at identifying through characteristic vibrational bands, such as C-H stretching modes in the 2780–2980 cm⁻¹ range, which correspond to the polymer backbones of materials like polyethylene and polystyrene. This technique has been applied to quantify microplastics in deionized and marine waters, achieving detection limits down to 1 μm with minimal interference from aqueous media. Recent advancements in the 2020s have extended Raman imaging to nanoplastics, using stimulated Raman scattering for 3D mapping of particles smaller than 1 μm, addressing challenges like weak scattering signals through signal enhancement algorithms and high-resolution optics. In air quality monitoring, Raman spectroscopy provides insights into aerosol composition by analyzing individual particles, identifying components such as sulfates via sulfate ion peaks around 980 cm⁻¹ and distinguishing them from organics or soot. Single-particle Raman analysis has been used to characterize atmospheric aerosols in real-time, revealing mixing states and phase transitions that influence climate forcing and air pollution dynamics. Surface-enhanced variants further improve sensitivity for trace organics in ambient aerosols, enabling detection of pollutants at concentrations relevant to urban and remote environments. For soil and sediment analysis, Raman spectroscopy characterizes mineralogy and organic contaminants by probing silicate vibrations for minerals like (464 cm⁻¹) and kaolinite, alongside C-H and C=O bands for hydrocarbons and pesticides. It facilitates the speciation of anthropogenic phases in clay-rich estuarine sediments, distinguishing natural organics from pollutants without disrupting the sample matrix. This approach has been employed to study mineral-organic interfaces, elucidating carbon sequestration processes and contaminant mobility in terrestrial ecosystems. Remote sensing applications leverage standoff Raman spectroscopy for the detection of hazardous materials at distances up to several meters, using deep-ultraviolet excitation to minimize fluorescence and enhance safety in field scenarios. Systems combining Raman with have identified explosives and chemical agents on surfaces, providing rapid, non-contact identification critical for environmental hazard response. Emerging in situ aquatic probes based on Raman spectroscopy enable real-time monitoring of algae blooms by classifying harmful cyanobacterial species through pigment and lipid signatures, such as phycocyanin bands at 1580–1650 cm⁻¹, aiding in the prediction of eutrophication events. For oil spills, underwater Raman microscopy detects petroleum hydrocarbons via aromatic C-H stretches around 3000 cm⁻¹, supporting rapid assessment of spill extent and degradation in marine environments.

Advanced Techniques

Micro-Raman Spectroscopy

Micro-Raman spectroscopy, often implemented through confocal Raman microscopy, adapts conventional for high-spatial-resolution analysis by integrating it with confocal optical microscopy. The setup typically couples a laser excitation source, such as a 514.5 nm or 785 nm line, to a microscope objective with high numerical aperture (N.A., e.g., 0.75–0.95) for focusing the beam onto the sample. Raman-scattered light is collected through the same objective and passed through a confocal pinhole (typically 70–100 μm) before entering a spectrometer equipped with a notch filter to block Rayleigh scattering and a charge-coupled device (CCD) detector for signal acquisition. This configuration achieves laser spot sizes down to approximately 1 μm laterally, enabling localized vibrational analysis of microscopic regions without extensive sample preparation. The technique was pioneered by Delhaye and Dhamelincourt in 1975, who developed the first confocal Raman microscope to extend capabilities to small particles. Spatial resolution in micro-Raman spectroscopy is fundamentally constrained by optical diffraction, with the minimum spot diameter given by approximately d = 0.61 \lambda / \text{NA}, where \lambda is the excitation wavelength; for example, this yields ~0.44 μm at 514.5 nm and N.A. 0.75. The confocal pinhole enhances axial resolution by spatially filtering out-of-focus contributions, reducing the depth of field to 1–2 μm and improving contrast by rejecting background fluorescence or stray light from underlying layers. These improvements surpass standard wide-field Raman setups, where diffraction limits both lateral and axial resolutions without such filtering, allowing micro-Raman to resolve features at the diffraction limit while minimizing interference. A key application of micro-Raman spectroscopy lies in mapping chemical heterogeneity across samples, such as identifying phase distributions in polycrystalline materials or compositional variations in thin films, by raster-scanning the focused spot in two dimensions. This provides spatially resolved Raman spectra that reveal molecular fingerprints at micron scales, facilitating non-destructive characterization of complex structures. The primary advantages include the capability for 3D chemical profiling through sequential depth scanning (z-stacking), which reconstructs volumetric distributions of analytes with sub-micron precision, offering insights into layered or heterogeneous systems. However, challenges persist, particularly sample heating due to high laser irradiance (e.g., >10 mW focused to ~1 μm spot), which can induce thermal degradation or changes; this is mitigated by using longer wavelengths (e.g., 785 nm) or pulsed to limit .

Polarized Raman Spectroscopy

Polarized Raman spectroscopy extends conventional Raman techniques by incorporating polarization control of both the incident light and the collected scattered light, enabling the probing of and orientation in samples. This method leverages the anisotropic nature of , where the intensity of scattered light varies with the relative orientations of the vectors of the incident and scattered photons. The depolarization ratio, defined as ρ = I_⊥ / I_∥, where I_⊥ and I_∥ are the intensities of the perpendicular and parallel components of the scattered light relative to the incident , serves as a key parameter for classifying vibrational mode symmetries. For totally symmetric modes (e.g., in point groups like C_{2v}), ρ approaches 0, indicating isotropic scattering, while depolarized modes (e.g., or E symmetries) exhibit ρ ≈ 3/4, reflecting anisotropic changes. The experimental setup for polarized Raman spectroscopy typically involves inserting polarizers in both the and collection optical paths of a standard Raman spectrometer. A in the excitation arm ensures linearly polarized incident light, often along a defined direction (e.g., vertical or horizontal), while an analyzer in the collection path selects either (∥) or (⊥) components of the scattered light. Configurations are described using Porto notation, such as Z(XX)Ȳ for backscattering with polarizations or Z(XY)Ȳ for , where the first and last letters denote directions, and the middle letters indicate directions. Half-wave plates may be used to rotate the incident without altering the light's , allowing systematic variation of angles for studies. This setup reveals angular dependencies in , particularly in oriented or crystalline samples. Raman tensors, which are 3×3 matrices describing the polarizability changes associated with a vibrational , provide a mathematical framework for interpreting data and determining . Each tensor element (e.g., α_{xx}, α_{yy}, α_{zz}) quantifies how the oscillates along principal axes during the vibration, with isotropic tensors (trace-dominant, off-diagonal elements zero) corresponding to symmetric and anisotropic tensors (off-diagonal or unequal diagonal elements) to asymmetric ones. The ratio can be derived from tensor invariants: ρ = 3γ² / (45ā² + 4γ²), where ā is the mean and γ² the , yielding ρ < 0.75 for polarized (symmetric) bands and ρ = 0.75 for fully (antisymmetric) bands. In oriented systems, tensor orientation relative to the sample axes dictates intensity variations, enabling assignment (e.g., A_g vs. B_g in D_{4h} crystals). Excitation profile analysis, varying wavelength while monitoring , can tie to weak effects but primarily highlights without deep electronic coupling. Applications of polarized Raman spectroscopy are prominent in characterizing anisotropic materials, such as determining crystal orientation through intensity angular profiles. For instance, in crystals, polarization-dependent spectra distinguish domain orientations by varying Raman peak intensities with rotation angles, aiding ferroelectric device fabrication. In polymers, it maps molecular and stress distribution; polarized measurements on extruded poly() composites reveal uniaxial orientation via the angular dependence of specific bands, quantifying Hermans orientation factors up to 0.8. Liquid crystals benefit from this technique for assessing , as seen in nematic phases where perpendicular configurations show enhanced perpendicular , enabling monitoring of phase transitions and anchoring effects on substrates.

Resonance Raman Spectroscopy

Resonance Raman spectroscopy is a variant of Raman scattering where the excitation laser wavelength is tuned to closely match or resonate with an electronic absorption transition of the analyte, resulting in a dramatic increase in the intensity of specific vibrational bands associated with that electronic state. This resonance condition promotes the molecule to a real electronic excited state, enabling rapid virtual or real transitions that couple electronic and vibrational motions, thereby selectively enhancing the Raman cross-section for modes involved in the Franck-Condon active distortions. Unlike non-resonant Raman, this process allows detection of weak signals from dilute samples or transient species, with enhancement factors typically ranging from $10^2 to $10^6 relative to off-resonance conditions. The enhancement mechanism is theoretically described by the Kramers-Heisenberg-Dirac dispersion formula, which expresses the Raman polarizability tensor. In the preresonance regime—where the energy approaches but does not exactly match the electronic transition—the dominant contribution arises from Albrecht's A-term, reflecting vibronic via Franck-Condon factors. This term approximates the as: \alpha_{fi} \approx \sum_e \frac{\langle \chi_f | \chi_e \rangle \langle \chi_e | \chi_i \rangle \langle \mu_e | \mu_g \rangle^2}{E_e - E_i - \hbar\omega_L + i\Gamma_E} Here, \chi denotes vibrational wavefunctions, \mu the transition dipole moments, E_e and E_i the energies of the excited and initial states, \omega_L the laser frequency, and \Gamma_E a damping factor accounting for excited-state lifetime broadening; the small denominator near resonance amplifies the intensity. This approximation simplifies calculations for near-resonant conditions, highlighting how distortions in the excited state selectively boost certain vibrational progressions. A key application lies in probing conjugated chromophores, such as those in proteins like and , where resonance excitation (e.g., at 413–418 nm for the Soret band) selectively enhances ring vibrations, revealing details on Fe-ligand bonds and peripheral substituents without interference from protein backbone modes. This selectivity arises from the strong vibronic coupling in extended \pi-conjugated systems, enabling structural insights into states and binding. For biomolecules lacking visible chromophores, (UV) excitation (e.g., 244–266 nm) is commonly employed to target aromatic residues like and , while shifting the Raman-shifted Stokes lines into the visible range to minimize interference from endogenous fluorophores. Despite these advantages, resonance Raman is limited by potential photodegradation of sensitive samples, particularly under UV or high-intensity visible excitation, which can induce covalent damage, crosslinking, or in biomolecules like proteins and nucleic acids, necessitating low powers and short exposures to preserve sample .

Variants

Spontaneous Raman Spectroscopy

Spontaneous Raman spectroscopy is the foundational technique in Raman analysis, relying on the incoherent, of monochromatic by molecular vibrations and rotations in the far field. This linear response occurs when incident photons interact with the sample, resulting in scattered shifted in energy by amounts corresponding to vibrational transitions. The process is characterized by an extremely low scattering cross-section, typically on the order of $10^{-30} cm² per molecule, which limits the efficiency of signal generation compared to other spectroscopic methods like . Standard experimental setups for spontaneous Raman spectroscopy utilize either dispersive or Fourier transform (FT) configurations. In dispersive systems, a continuous-wave illuminates the sample, and the scattered light is collected, filtered to remove , and dispersed by a onto a (CCD) detector for . FT-Raman systems, often employing near-infrared excitation to minimize , use an interferometer to encode the spectral information interferometrically, followed by to recover the spectrum; these are particularly advantageous for samples prone to . Both approaches enable high , typically down to 1 cm⁻¹, and are compatible with objectives for spatially resolved measurements. A key strength of spontaneous Raman spectroscopy lies in its ability to provide direct, quantitative insights into molecular and without altering the sample, as it is inherently non-destructive and label-free. This allows for the of , liquids, gases, and biological materials in their native states, yielding information on chemical bonds, polymorphism, and through and positions. However, the technique's primary limitation is the feeble signal due to the small cross-section, necessitating long acquisition times—often from seconds to hours per spectrum—to achieve sufficient signal-to-noise ratios, which can hinder real-time applications. As the baseline for Raman variants, spontaneous Raman serves as a reference for unmodified linear , where subsequent techniques introduce enhancements or nonlinear processes to overcome its constraints without fundamentally altering the underlying vibrational probing .

Enhanced Raman Spectroscopy

Enhanced Raman spectroscopy encompasses techniques that dramatically amplify Raman signals through localized enhancements, primarily via plasmonic nanostructures, enabling ultrasensitive detection far beyond conventional Raman methods. The two primary variants are (SERS) and tip-enhanced Raman spectroscopy (TERS), both leveraging surface plasmons on noble metals to achieve enhancement factors ranging from 10^6 to 10^14, depending on substrate design and positioning. These methods are particularly valuable for analyzing trace molecules where spontaneous Raman signals are too weak, such as in single-molecule studies. SERS, first observed in the 1970s, involves adsorbing s onto roughened metal surfaces or nanostructured substrates, typically silver (Ag) or (Au) nanoparticles or films, where localized surface s generate intense electromagnetic fields. The enhancement arises predominantly from the electromagnetic (EM) mechanism, in which the incident and scattered couple to plasmon resonances, amplifying the local field intensity by factors up to 10^8 in optimized "hot spots"—nanoscale junctions between particles where fields concentrate due to coupling effects. A secondary chemical enhancement mechanism contributes modestly (typically 10^1 to 10^3), involving charge-transfer resonances between the analyte and substrate that modify molecular . Overall, these combined effects enable detection limits down to attomolar concentrations for certain analytes. TERS extends SERS by integrating a plasmonic tip, often an Ag- or Au-coated (AFM) probe, to achieve sub-10 nm spatial resolution through near-field confinement. In gap-mode TERS, the is positioned in the narrow gap between the tip and a , exploiting highly confined plasmons for enhancement factors comparable to SERS while providing topographic and spectroscopic mapping at the nanoscale. This configuration minimizes far-field contributions, ensuring signal localization to ~1-10 nm volumes. Key applications of SERS and TERS include single-molecule detection, where stochastic blinking in spectra confirms the presence of individual analytes like Rhodamine 6G on colloids, and trace analysis of environmental pollutants or biomolecules at picomolar levels. In biomedical contexts, these techniques enable label-free identification of cancer biomarkers in cells. Recent advancements in the have focused on biocompatible SERS substrates, such as silica- or polymer-coated nanoparticles, to enable applications like real-time tumor monitoring in animal models without . These designs address near-field stability in physiological environments, achieving sustained enhancements for multiplexed biosensing.

Non-Linear and Coherent Raman Methods

Non-linear and coherent Raman methods represent advanced variants of Raman spectroscopy that leverage multi-photon interactions to generate enhanced signals, enabling faster and more sensitive imaging compared to spontaneous Raman techniques. These methods, primarily Coherent Anti-Stokes Raman Scattering (CARS) and Stimulated Raman Scattering (SRS), rely on third-order nonlinear optical processes where the interaction of multiple laser beams coherently drives molecular vibrations, producing directional, laser-like emission without the stochastic nature of spontaneous scattering. Developed in the late 20th century, these techniques have revolutionized label-free chemical imaging by providing high-speed, background-reduced detection of molecular fingerprints. Coherent Anti-Stokes Raman Scattering (CARS) is a four-wave mixing process involving a pump beam at frequency \omega_p, a Stokes beam at \omega_s, and often a probe beam (which can be the same as the pump), generating an anti-Stokes signal at \omega_{as} = 2\omega_p - \omega_s. When the frequency difference \omega_p - \omega_s matches a molecular vibrational frequency, the third-order nonlinear susceptibility \chi^{(3)} resonantly enhances the signal, combining both resonant (vibration-specific) and non-resonant background contributions. The process requires phase-matching for efficient signal generation, typically achieved through collinear beam geometry in microscopy setups, as pioneered by Zumbusch et al. in 1999 for three-dimensional sectioning. Early observations of CARS date back to Terhune and Maker in 1965, establishing it as a coherent analog to spontaneous Raman with significantly higher signal intensity. Stimulated Raman Scattering (SRS), in contrast, detects the energy transfer between a pump beam (\omega_p) and a Stokes beam (\omega_s) when \omega_p - \omega_s aligns with a vibrational mode, leading to either stimulated Raman gain (SRG) in the Stokes beam or stimulated Raman loss (SRL) in the pump beam as a depletion signal. This second-order process in terms of detection is purely resonant, avoiding the non-resonant background inherent in CARS, and scales linearly with analyte concentration for quantitative imaging. SRS microscopy, enabling video-rate acquisition, was advanced by Freudiger et al. in 2008 through modulation techniques like lock-in detection to isolate the weak SRL signal against laser noise. The foundational work on SRS traces to Woodbury and Ng in 1962, who observed the stimulated emission in liquids. These methods offer key advantages, including high speeds up to video rates (e.g., 20 frames per second), absence of , and inherent for label-free visualization of molecular distributions. In applications, and excel in live-cell dynamics, such as tracking in adipocytes or in neurodegenerative diseases, and in three-dimensional for intraoperative tumor margin detection. For instance, has enabled real-time monitoring of drug distribution in pharmaceuticals, while provides rapid 3D mapping of myelin sheaths in neural tissues. However, challenges persist, including the need for precise phase-matching to maximize , complex optical alignment of multiple beams, and sensitivity limitations from non-resonant backgrounds in or detection noise in . These hurdles often necessitate specialized or lasers and advanced .

Emerging Variants

Morphologically directed Raman spectroscopy (MDRS) integrates automated particle with Raman chemical to characterize the , , and of individual particles in complex mixtures, particularly subvisible particles in pharmaceutical formulations. This technique employs high-resolution optical to select particles based on , followed by targeted Raman spectroscopy for molecular identification, enabling the discrimination of polymorphs, contaminants, or aggregates without . Recent advancements include its with scanning (SEM) for enhanced morphological detail at the nanoscale, allowing simultaneous elemental mapping and vibrational fingerprinting in forensic and materials applications. For instance, in detection, SEM-Raman hybrids identify both inorganic and organic components on a single platform, improving accuracy in characterization. Spatially offset Raman spectroscopy (SORS) enables non-invasive detection of subsurface chemical compositions through turbid or opaque barriers by collecting Raman signals from points laterally displaced from the spot, thus minimizing surface . Developed for applications like pharmaceutical packaging integrity and analysis, SORS has evolved with micro-SORS variants that achieve sub-millimeter resolution for layered samples, such as reconstructing hidden text in historical manuscripts. In 2025, experimental protocols have quantified depth-sensing limits, establishing offsets up to several millimeters for biomedical uses like composition assessment . Handheld SORS devices now facilitate real-time subsurface probing, with signal depths scaling linearly with offset distance in diffusive media. Ultrafast Raman spectroscopy utilizes laser pulses to capture time-resolved vibrational dynamics on timescales, revealing transient molecular structures during chemical reactions or energy transfer processes. Impulsive stimulated Raman variants achieve broadband detection from THz to 3000 cm⁻¹, enabling observation of coherent vibrational wavepackets in photochemical intermediates. A 2024 advancement includes optimized algorithms for processing time-resolved coherent Raman data, reducing computational overhead while preserving spectral fidelity. These methods have quantified bond-breaking timescales in photoexcited molecules, with resolutions below 100 fs, advancing understanding of ultrafast processes in and . Integrations of (AI) and principles are transforming Raman data handling through (Findable, Accessible, Interoperable, Reusable) platforms and (ML) models for hyperspectral processing. In 2025, open repositories have aggregated large Raman datasets, enabling ML algorithms to automate peak assignment and in multidimensional spectra, with accuracy improvements up to 95% for complex mixtures. These digital tools facilitate collaborative analysis via standardized ontologies, accelerating discoveries in materials screening by predicting properties from spectral patterns without manual intervention. For hyperspectral Raman imaging, convolutional neural networks now classify features in real-time, supporting applications in and while ensuring data provenance through blockchain-linked metadata. Hybrid modalities combining Raman spectroscopy with (MS) or (MRI) provide multimodal insights by merging vibrational, mass-to-charge, and spatial data for comprehensive sample characterization. Raman-MS integrations, such as systems supporting both techniques, enable sequential molecular identification and fragmentation analysis in single-particle studies, enhancing for organics in environmental samples. In nanoparticle-based probes, triple-modality setups incorporating Raman, photoacoustic, and MRI signals allow tracking of tumor-targeted agents, with Raman providing label-free confirmation of chemical payload integrity. These hybrids have demonstrated sub-cellular resolution in preclinical , where Raman spectra correlate with MS-derived molecular weights to validate therapeutic delivery.

References

  1. [1]
    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.
  2. [2]
    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 ...
  3. [3]
    Artifacts and Anomalies in Raman Spectroscopy: A Review on ...
    Oct 8, 2024 · Sir C.V. Raman discovered the Raman effect in 1928, which forms the basic principle of Raman spectroscopy based on the inelastic scattering of ...
  4. [4]
  5. [5]
    [PDF] Raman Spectroscopy
    Raman Spectroscopy involves the study of transitions between quantum levels of molecules and matter induced by the inelastic scattering of light.
  6. [6]
    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.
  7. [7]
    The role of Raman spectroscopy in biopharmaceuticals from ... - NIH
    Raman spectroscopy offers many benefits and is used successfully in bioprocessing from single-cell analysis to cGMP process control.
  8. [8]
    Novel aspects of Raman spectroscopy in skin research - PMC - NIH
    2. HISTORY OF RAMAN SPECTROSCOPY. The principle of Raman spectroscopy is based on the Raman effect which has been discovered more than 90 years ago. The ...
  9. [9]
    A New Radiation: C.V. Raman and the Dawn of Quantum ...
    Mar 3, 2025 · The Raman Effect. The history of Raman spectroscopy begins with the discovery of the "scattering effect" by C.V. Raman and Kariamanikam ...Missing: definition | Show results with:definition
  10. [10]
    (PDF) The Discovery of the Raman Effect and Early Applications of ...
    Aug 9, 2025 · PDF | On Apr 1, 2011, Rajinder Singh published The Discovery of the Raman Effect and Early Applications of Infrared and Raman Spectroscopy ...
  11. [11]
    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.
  12. [12]
    This Month in Physics History | American Physical Society
    In February 1928 Raman observed that the scattered light was polarized, which distinguished the new scattering effect from fluorescence. He and colleague K.S. ...
  13. [13]
    A New Type of Secondary Radiation | Nature
    inelastic insight​​ Motivated by Arthur Compton's observation that X-rays could lose energy when scattered inelastically by electrons (the ' ...
  14. [14]
    [PDF] Sir Chandrasekhara V. Raman - Nobel Lecture
    The scattering of light in transparent fluids is extremely feeble, much weaker in fact than the Tyndall effect usually ob-.
  15. [15]
    C. V. Raman and the Discovery of the Raman Effect
    In 1928 the Indian physicist C. V. Raman (1888-1970) discovered the effect named after him virtually simultaneously with the Russian physicists G. S. Lands.
  16. [16]
    Principles of Raman spectroscopy (4) History of ... - JASCO Global
    Dec 9, 2021 · After the principle of Raman scattering was developed by CV Raman et al., the main light source used was a mercury lamp.
  17. [17]
    An Introduction to Raman Spectroscopy - Wiley Analytical Science
    Jun 30, 2014 · The genesis of Raman spectroscopy was in the first quarter of the 20th century when the scattering of monochromatic radiation with change of ...
  18. [18]
    [PDF] Monograph: Introduction to Raman spectroscopy - Metrohm
    In the 1960s, when the first lasers be came available, Raman spectroscopy be came a popular spectroscopic technique for a wide range of chemical ana lyses.
  19. [19]
    Discover 50 years of Raman innovation by HORIBA
    Raman spectroscopy history started in the 1920s. Raman spectroscopy was named after Sir Chandrasekhara Venkata Raman (7 November 1888 – 21 November 1970), an ...
  20. [20]
    Evolution of Instrumentation for Detection of the Raman Effect as ...
    Aug 6, 2025 · The CCD camera was applied first in the field of Raman spectroscopy for solids in 1987 for characterizing ultrathin organized layers of organic ...
  21. [21]
    What is Raman Spectroscopy? - HORIBA
    In 1990 the holographic notch filters were demonstrated to provide superior laser rejection so that a Raman microscope could be built on a single stage ...
  22. [22]
    Fourier transform Raman spectroscopy - ACS Publications
    Article November 1, 1986. Fourier transform Raman spectroscopy. Click to copy article linkArticle link copied! D. Bruce. Chase. ACS Legacy Archive. Open PDF ...
  23. [23]
    Portable Raman Spectrometers: How Small Can They Get?
    Jun 1, 2023 · Portable Raman spectrometers have become smaller over the last 20 years, while their performance has increased.
  24. [24]
    Advances in real‐time fiber‐optic Raman spectroscopy for early ...
    Sep 25, 2020 · This article reviews the cutting-edge clinical Raman instrumentation, developments in fiber-optic Raman endoscopic probes, spectral ...
  25. [25]
    AI-Powered Raman Spectroscopy Signals New Era for Drug ...
    Jun 4, 2025 · Deep learning algorithms improve spectral analysis, automatically identifying complex patterns in noisy Raman data and reducing the need for ...
  26. [26]
    The quantum theory of the emission and absorption of radiation
    The quantum theory of dispersion. P. A. M. Dirac, Proceedings A, 1927. On the theory of quantum mechanics. Paul Adrien Maurice Dirac, Proceedings A, 1926. The ...
  27. [27]
    Raman: Theory - Chemistry LibreTexts
    Jan 29, 2023 · Classical theory · is termed the polarizability tensor. It is a second-rank tensor with all the components in the unit of CV-1m2. Typically, ...Introduction · Classical theory · Quantum mechanical treatment
  28. [28]
    Unified Treatment of Fluorescence and Raman Scattering ...
    Dec 6, 2004 · The model studies surface-enhanced Raman and fluorescence, focusing on electromagnetic effects and molecular dynamics, treating them equally. ...
  29. [29]
    Raman Spectroscopy
    Raman spectroscopy uses molecular vibrations causing a change in polarizability, where molecules are irradiated by photons and emit Raman photons.
  30. [30]
    Stokes and anti-Stokes scattering - DoITPoMS
    The difference in intensity of the Stokes and anti-Stokes components is due to the different number of molecules in each state initially.
  31. [31]
    Raman Thermometry: Understanding the Mathematics to Better ...
    Dec 1, 2019 · Raman measurements can be used to infer the temperature of a material based on the Stokes and anti-Stokes band signal strengths, and by applying a predefined ...
  32. [32]
    Power of Infrared and Raman Spectroscopies to Characterize Metal ...
    Dec 14, 2020 · In the near infrared (14000–4000 cm–1) overtones and combination modes are mainly detected. The low-energy far-infrared (ca.
  33. [33]
    Criteria for choosing a Raman spectrometer. - InPhotonics
    Spectral Range: The Raman shift range is specified in wavenumbers (cm-1). Most spectrographs cover at least the "fingerprint region" from 400 - 1800 cm-1 ...
  34. [34]
    High-temperature and high-pressure Raman spectra of Fo 89 Fa 11 ...
    Oct 1, 2021 · All the Raman-active modes systematically shift to lower frequencies at elevated temperatures, whereas to higher frequencies with increasing ...
  35. [35]
    Effects of pressure and isotopic substitution on the Raman spectrum ...
    Apr 15, 1990 · We report measurements of the effects of high pressure and O-isotope substitution on the Raman spectrum of α- F e 2 O 3.
  36. [36]
    [PDF] Relative Intensity Correction of Raman Spectrometers
    checked by measuring the Raman shift position of the silicon phonon (520 cm 1) and also the spectrum of a number of. Raman shift standards proposed by the ASTM.
  37. [37]
    Verification of Si wafer first‐order phonon peaks for reliable ...
    Nov 27, 2023 · Although the first-order phonon peak of Si wafers at ⁓520 cm−1 is generally used as a calibrant of Raman microscopes, not only is it unclear ...
  38. [38]
    Lasers for Raman Spectroscopy - RP Photonics
    There are lasers which are specifically suitable for application Raman spectroscopy. We discuss the requirements and types of lasers.<|control11|><|separator|>
  39. [39]
    Laser Raman Spectroscopy with Different Excitation Sources and ...
    Aug 3, 2015 · In this paper, we report the use of a dispersive Laser Raman system operable with three different excitation sources (Argon-ion, He-Ne, and Diode lasers)
  40. [40]
    What Lasers are Best for Raman Spectroscopy? - AZoOptics
    Oct 4, 2023 · 785 nm lasers are commonly used. 1064 nm Nd:YAG lasers are also used. 532 nm visible lasers are used for SERS. The best laser depends on the ...
  41. [41]
    Laser wavelength selection in Raman spectroscopy - RSC Publishing
    Apr 7, 2025 · In this article we give a full presentation and discussion of the various factors to be considered when choosing a laser wavelength for Raman spectroscopy.
  42. [42]
    Safety Considerations for Sample Analysis Using a Near-Infrared ...
    Raman spectroscopy is often considered a nondestructive analytical technique; however, this is not always the case. The 300-mW 785-nm near-infrared (NIR) ...
  43. [43]
    Safety considerations for sample analysis using a near-infrared (785 ...
    The 785-nm laser can burn samples, especially irreplaceable or hazardous ones. Darker colors are more problematic. Safety guidelines are presented.Missing: excitation | Show results with:excitation
  44. [44]
    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.
  45. [45]
    Resolution of Monochromators and Spectrographs - Newport
    The resolution can be theoretically determined by multiplying the reciprocal dispersion of the grating by the slit width. The monochromator bandpass with a 1200 ...
  46. [46]
    Spectral Resolution in Raman Spectroscopy - Edinburgh Instruments
    Five main factors determine the spectral resolution achievable: slit size, diffraction grating, spectrometer focal length, detector, and the excitation laser.
  47. [47]
    Detectors for Raman Spectroscopy: CCDs, EMCCDs, and InGaAs ...
    Apr 26, 2023 · CCDs are the most commonly used detector in Raman spectroscopy. BI CCDs offer higher QE than FI CCDs, but are prone to etaloning, which can be ...
  48. [48]
    Raman spectroscopy - Wikipedia
    Raman spectroscopy is commonly used in chemistry to provide a structural fingerprint by which molecules can be identified.Resonance Raman · Stimulated Raman spectroscopy · Surface-enhanced Raman
  49. [49]
    Raman spectroscopy - Hamamatsu Photonics
    InGaAs linear image sensor G14237-512WA. This is an InGaAs linear image sensor designed for Raman spectroscopy measurement using a 1064 nm laser. Designed ...
  50. [50]
    White Paper: Understanding the Advantages of FT-Raman.
    Fellgett's advantage, also known as the multiplex advantage, arises from how data is distributed in an interferogram. In Figure 1, we illustrate interferogram/ ...
  51. [51]
    Understanding Raman Spectrometer Parameters
    Just as a longer exposure gives you a better picture in dim light, you get a better spectrum from a weak Raman scatterer. Spectroscopy is static. As such ...Missing: early | Show results with:early
  52. [52]
    FT-IR Spectroscopy - Newport
    Multiplex (Fellgett) Advantage​​ In a dispersive spectrometer, wavenumbers are observed sequentially, as the grating is scanned. In an FT-IR spectrometer, all ...
  53. [53]
    [PDF] RAM II FT-Raman Module
    Mar 8, 2022 · The ´hybrid` platform enables the use of both FT-Raman microscopy at 1064nm and dispersive Raman microscopy in the visible. MultiRAM. Bruker´s ...
  54. [54]
  55. [55]
    Raman spectroscopy: techniques and applications in the life sciences
    This review is intended to be a starting resource for scientists new to Raman spectroscopy, providing theoretical background and practical examples as the ...
  56. [56]
    Key Elements of Confocal Raman Microscopy for High-Resolution ...
    The pinhole plays a crucial role for the depth resolution in confocal microscopy. However, the physical size of the pinhole (P physical) cannot easily be ...
  57. [57]
    Basic Principles of Raman Scattering and Spectroscopy
    An achromatic lens then focuses the light onto the sample. The light that scatters off the sample hits the second prism mirror, which deflects it into a beam ...
  58. [58]
    [PDF] Raman Spectroscopy: Successes, Challenges, and Pitfalls
    Dec 9, 2020 · Raman is applicable to solids (amorphous to crystalline, mineral or organic), liquids, and gases (Fig. 5). In transparent samples, analyses can ...
  59. [59]
  60. [60]
  61. [61]
    A Comprehensive Review on Raman Spectroscopy Applications
    Raman spectroscopy is a very powerful tool for material analysis, allowing for exploring the properties of a wide range of different materials.
  62. [62]
  63. [63]
    Raman spectroscopic study of diamond and graphite in ureilites and ...
    Nov 8, 2012 · It is most likely that this peak is due to lonsdaleite that is hexagonal diamond and one of the polymorphs of diamond and graphite. It has ...Missing: polymorphism | Show results with:polymorphism
  64. [64]
    Stress, Strain, and Raman Spectroscopy
    Sep 1, 2019 · When stress is applied to an object, it can produce strain. Strain can be detected through changes in peak position and bandwidth in Raman ...
  65. [65]
    In situ Raman spectroscopic studies of CO2 reduction reactions
    This article reviews recent advancements in the application of in situ Raman spectroscopy in eCO 2 RR research, focusing on its critical role in monitoring ...
  66. [66]
    Direct estimation of polymer crystallinity with Raman spectroscopy ...
    Raman spectroscopy had also been used to estimate percent crystallinity of polymers such as polystyrene [9], polylactide [10], polypropylene [15], polyethylene ...
  67. [67]
    Applications of Raman Spectroscopy in Art and Archaeology - Rousaki
    Dec 2, 2024 · Raman spectroscopy identified tenorite (CuO) as the pigment for the black finishing appearance, ascribing the patina to an intentional artist's ...
  68. [68]
    [PDF] Raman spectroscopy in cultural heritage: Background paper
    Jun 4, 2015 · In the last 30 years Raman spectroscopy has become an essential experimental tool for the analysis of a wide variety of artists' materials,.
  69. [69]
    In situ Raman spectroscopy for cultural heritage studies
    Jun 3, 2021 · In cultural heritage studies, the preferable Raman systems for in situ analysis are mobile instruments or portable ones with long fibre-optics ...Missing: Portable analysis
  70. [70]
    Pigments in Ancient Manuscripts and Paintings Brought to Life by ...
    Oct 17, 2022 · Raman spectroscopy led to the identification of cinnabar, red lead, haematite, goethite, lead white, verdigris, azurite and caput mortuum.
  71. [71]
    Raman spectroscopic library of medieval pigments collected with ...
    This study created a Raman library of 32 medieval pigment spectra using five wavelengths to help identify pigments in manuscripts and plan laser wavelength use.Missing: ancient | Show results with:ancient
  72. [72]
    A kinetic model of red vermilion darkening from Raman spectra
    Among the principal pigments used for decorating artifacts or paintings, mercury sulfide HgS, known as cinnabar, assumed a prominent role as red color for many ...Missing: peak | Show results with:peak
  73. [73]
    Recognition of fake paintings of the 20th-century Russian avant ...
    Sirro et al. utilized Raman spectroscopy on a counterfeit case of fake paintings of the 20th-century Russian avant-garde [64]. The study showed that Raman was ...
  74. [74]
    Can Raman microscopy reveal art forgeries? - HORIBA
    Raman spectroscopy is a non-destructive chemical analysis technique that uses inelastic light scattering to identify a molecule's unique fingerprint. It ...
  75. [75]
    Cultural Heritage Meets the Art Forensics Enigma - SpringerLink
    Oct 14, 2022 · Cultural Heritage Meets the Art Forensics Enigma: Raman Spectroscopic Authentication and the Exposure of Fakes and Forgeries in Art Works.Missing: spectroscopy | Show results with:spectroscopy
  76. [76]
    [PDF] Raman study of different crystalline forms of PbCrO4 and PbCr1 ...
    Jun 14, 2014 · Noninvasive Raman spectroscopy has been successfully exploited to study in situ the yellow palette of three Van Gogh paintings. The results, ...
  77. [77]
    Identification of pigments used in a Qajar manuscript from Iran by ...
    Mar 3, 2022 · Raman spectroscopy was used to identify pigments based on molecular vibrations. Raman spectra of all inks and pigments, except for the golden ...
  78. [78]
    A Review of Non-Destructive Raman Spectroscopy and ... - MDPI
    Nov 12, 2024 · Furthermore, portable Raman systems make it possible to conduct fieldwork in heritage sites without transporting samples to laboratories, ...<|control11|><|separator|>
  79. [79]
    The examination of Van Gogh's chrome yellow pigments in 'Field ...
    Dec 5, 2019 · Alternatively, Fourier Transform Infrared Spectroscopy (FTIR) or Raman-analysis can be used to differentiate between the types of chrome yellow ...
  80. [80]
    Raman spectroscopy: techniques and applications in the life sciences
    This review briefly presents the physical origin of Raman scattering, explaining the key classical and quantum mechanical concepts. Variations of the Raman ...
  81. [81]
    None
    ### Summary of Raman Microscopy from https://www.chem.uci.edu/~dmitryf/manuals/Fundamentals/Raman%20microscopy.pdf
  82. [82]
    Confocal Raman Microscopy
    ### Summary of Micro-Raman Spectroscopy from Confocal Raman Microscopy (Springer, 2018)
  83. [83]
    [PDF] Recent Developments in Multichannel Raman Microprobing
    This new progress includes: confocal Raman microanalysis that provides better spatial resolution and background rejection, confocal Raman mapping that re-.Missing: challenges | Show results with:challenges
  84. [84]
    None
    Summary of each segment:
  85. [85]
    [PDF] Polarised Raman Spectroscopy - Edinburgh Instruments
    The RM5 Raman Microscope can be equipped with fully computer-controlled polarisation optics to allow the user to easily collect and analyse spectra and maps in ...
  86. [86]
    Raman Tensors and their application in structural studies of ...
    The Raman tensor associated with a Raman band describes how the polarizability of the molecule oscillates with the molecular normal mode of vibration. Methods ...
  87. [87]
    Resonance Raman Spectroscopy - an overview | ScienceDirect Topics
    Resonance Raman spectroscopy (RRS) is a technique where a laser is tuned to match a molecule's electronic transition, enhancing vibrational scattering. It is a ...
  88. [88]
    Exploring Resonance Raman Spectroscopy
    Dec 1, 2018 · Resonance Raman spectroscopy cannot be described properly without invoking quantum mechanics and some mathematically complicated concepts in ...
  89. [89]
    Electronic Preresonance Stimulated Raman Scattering Microscopy
    Jul 12, 2018 · The theory of preresonance Raman was first worked out by Andreas C. Albrecht, who derived the Albrecht A-term preresonance approximation ...Author Information · Biographies · References
  90. [90]
    Resonance Raman Interrogation of the Consequences of Heme ...
    Resonance Raman spectroscopy has been used as an important tool for the study of heme proteins giving detailed structural information on the heme macrocycle ...Missing: systems | Show results with:systems
  91. [91]
    Resonance Raman Spectroscopy - an overview | ScienceDirect Topics
    The rR technique involves exciting a molecule using a laser having a frequency matching a given electronic transition and recording the resulting light- ...
  92. [92]
    Advantages of UV Raman - Wasatch Photonics
    Avoiding the fluorescence window​​ The efficiency of Raman scattering scales inversely with excitation wavelength to the fourth power (1/λex4), which in itself ...
  93. [93]
    Raman Crystallography of RNA - PMC - NIH
    However, resonance Raman has a number of limitations, especially for RNA. ... photodegradation of the RNA but still gives acceptable Raman scattering.
  94. [94]
    [PDF] Coherent Anti-Stokes Raman Scattering Microscopy and Its ...
    Dec 17, 2020 · The cross-section of Raman scattering is approximately 10-30 cm2 per molecule, which is 1011 times smaller than the corresponding cross ...Missing: fundamentals cm²
  95. [95]
    [PDF] Modern Raman Spectroscopy – A Practical Approach
    ... 1928 by Raman and. Krishnan [2]. Since then, the phenomenon has been referred to as Raman spectroscopy. In the original experiment sunlight ... benzene are shown ...
  96. [96]
    [PDF] Raman Spectroscopy: A Non-Destructive and On-Site Tool for ...
    Apr 20, 2012 · The setup is equipped with 12 different laser excitation wavelength, provided by: a 532 nm diode laser (Ventus LP 532), a Spectra Physics 632.8 ...
  97. [97]
    [PDF] Raman Spectroscopy and Microscopy of Individual Cells ... - OSTI
    Aug 17, 2008 · There are several advantages of. Raman spectroscopy: it requires no special dyes or specific excitation wavelengths, is mostly non-destructive ...
  98. [98]
    [PDF] Raman spectroscopy for unlabelled detection and quantification of ...
    As it is non-destructive, Raman spectroscopy can be combined with other ... spontaneous Raman spectroscopy to quantitative imaging or in vivo measurements, as.
  99. [99]
    [PDF] Spontaneous Compressive Raman technology: developments and ...
    Jun 30, 2021 · Raman scattering. The spontaneous Raman scattering cross-section is of the order of 10-30 cm2. This is to compare with fluorescence cross ...Missing: fundamentals cm²
  100. [100]
    Tip-Enhanced Raman Spectroscopy: Technique and Recent ...
    This review discusses a relatively new technique for optical nanoimaging at visible wavelength, known as tip-enhanced Raman spectroscopy (TERS).Metal for Enhanced... · Technique of TERS · Advanced Technologies in TERS
  101. [101]
    Surface Enhanced Raman Scattering Enhancement Factors
    This paper presents an in-depth study of Surface Enhanced Raman Scattering (SERS) enhancement factors (EFs) and cross-sections, including several issues ...
  102. [102]
    Single-Molecule Chemistry with Surface- and Tip-Enhanced Raman ...
    SERS and TERS use plasmonically enhanced Raman scattering to characterize the chemical information on single molecules.
  103. [103]
    Electromagnetic theories of surface-enhanced Raman spectroscopy
    Surface-enhanced Raman spectroscopy (SERS) and related spectroscopies are powered primarily by the concentration of the electromagnetic (EM) fields ...
  104. [104]
    A Review on Surface-Enhanced Raman Scattering - PMC - NIH
    Then, the measurement of the enhancement and the key factors that determine it (the materials, the hot spots, and the analyte-surface distance) are discussed. A ...
  105. [105]
    Tip-enhanced Raman spectroscopy for nanoscale probing of ...
    Nov 4, 2020 · Tip-enhanced Raman spectroscopy (TERS) is a near-field spectroscopic technique that combines the high spatial resolution of scanning probe ...
  106. [106]
    Tip-enhanced Raman spectroscopy: principles and applications
    Jul 1, 2015 · This review provides a detailed overview of the state of the art in tip-enhanced Raman spectroscopy (TERS) and focuses on its applications ...<|control11|><|separator|>
  107. [107]
    Recent development of surface-enhanced Raman scattering for ...
    May 6, 2023 · Recent developments of SERS substrates and their applications in biomolecular detection (SARS-CoV-2 virus, tumor etc.), biological imaging and pesticide ...Missing: 2020s | Show results with:2020s
  108. [108]
    In vivo surface-enhanced Raman scattering techniques - NIH
    Feb 11, 2025 · This review provides a comprehensive overview of SERS-based applications in vivo, focusing on key aspects such as the design considerations for SERS nanoprobes.Missing: 2020s | Show results with:2020s
  109. [109]
    Coherent Raman spectroscopy: Quo vadis? - ScienceDirect.com
    This paper aims to provide a holistic, forward-looking analysis of the two most widely applied CRS techniques, ie CARS and SRS, exploring their historical ...
  110. [110]
    Coherent Anti-Stokes Raman Scattering Microscopy and Its ...
    Dec 16, 2020 · Coherent anti-Stokes Raman scattering (CARS) microscopy can provide high resolution, high speed, high sensitivity, and non-invasive imaging of specific ...Introduction · CARS Principle · CARS Microscopy · Applications of CARS...Missing: paper | Show results with:paper
  111. [111]
  112. [112]
  113. [113]
    Stimulated Raman Scattering Microscopy: A Review - MDPI
    Stimulated Raman scattering (SRS) microscopy is a high-speed imaging modality based on molecular vibrations, producing chemical maps in living systems.
  114. [114]
  115. [115]
    Morphologically-Directed Raman Spectroscopy as an Analytical ...
    Nov 22, 2023 · Our method was applied to characterize particles formed in stressed TPs and was able to chemically identify individual particles using Raman spectroscopy.
  116. [116]
    Integrating Raman Spectroscopy with SEM for Advanced Materials ...
    Dec 13, 2024 · The RISE system is intended to give a thorough and complementary study of materials by combining the capabilities of Raman spectroscopy and SEM.
  117. [117]
    Combining Forensic SEM and Raman for Gunshot Residue Detection
    By integrating Raman with SEM, investigators can simultaneously detect both the metallic and organic components of GSR on a single platform.
  118. [118]
    [PDF] Advanced Microspatially Offset Raman Spectroscopy for ... - ePubs
    This paper explores the possibility of using micro-spatially offset Raman spectroscopy (micro-SORS) imaging to reconstruct non-invasively letters and ...
  119. [119]
    [PDF] Raman spectroscopy advancements in heritage science
    Micro-Spatially Offset Raman Spectroscopy (micro-SORS) is an advanced Raman technique that allows the non-destructive analysis of inner portions of cultural ...
  120. [120]
    Experimental method to assess depth sensing limits of inelastic ...
    Sep 18, 2025 · Aim: We aimed at developing an experimental method to establish a relationship between spatial offset in SORS and sampling depth. The technique ...
  121. [121]
    [PDF] Spatially offset Raman spectroscopy toward in vivo assessment of ...
    Jun 12, 2024 · In this study, we demonstrate the diagnostic utility of a hand- held SORS device for evaluating the chemical composition of the adipose tissue.<|separator|>
  122. [122]
    Tracking Ultrafast Structural Dynamics by Time-Domain Raman ...
    Jun 7, 2021 · Time-resolved impulsive stimulated Raman spectroscopy now realizes high sensitivity and a wide detection frequency window from THz to 3000 cm –1.
  123. [123]
    An ultrafast algorithm for ultrafast time-resolved coherent Raman ...
    Jan 4, 2025 · Time-resolved coherent Raman spectroscopy (CRS) is a powerful non-linear optical technique for quantitative, in-situ analysis of chemically ...
  124. [124]
    Review Towards understanding and controlling ultrafast dynamics in ...
    Nov 1, 2023 · In this contribution, we discuss recent experimental and theoretical developments seeking to understand and control the ultrafast dynamics associated with ...
  125. [125]
    Artificial Intelligence-Powered Raman Spectroscopy through Open ...
    Oct 27, 2025 · Raman spectra may be experimentally acquired, simulated through computational models, generated using different artificial intelligence (AI) ...
  126. [126]
    Open and FAIR Raman spectroscopy. Paving the way for artificial ...
    May 28, 2025 · We explore the current landscape of digital tools and open resources in Raman spectroscopy, highlight existing solutions as well as critical gaps.Missing: hyperspectral | Show results with:hyperspectral
  127. [127]
    [PDF] Development of a Dual Wavelength Microchip Laser for NASA's ...
    The RAMS laser system successfully demonstrated its ability to meet the requirements for both Raman Spectroscopy (RS) and Laser Desorption Mass Spectrometry ( ...
  128. [128]
    The Integration of Raman Spectrometers with Other Spectroscopic ...
    Modern systems offer integrated platforms that manage multiple instruments, coordinate acquisitions and calibration routines. The most important thing is that ...Why Integrate Raman... · 1. Raman-Infrared (ir)... · 2. Raman-Uv-Vis Spectroscopy...Missing: dispersive | Show results with:dispersive
  129. [129]
    Emergence of Raman Spectroscopy as a Probing Tool for ...
    Mar 5, 2023 · A unique triple-modality MPR nanoparticle has been reported, which stands for magnetic resonance imaging, photoacoustic imaging, and Raman ...<|separator|>