Fact-checked by Grok 2 weeks ago

Raman microscope

A Raman microscope is an analytical instrument that combines the principles of with optical microscopy to enable non-destructive chemical analysis and imaging of samples at microscopic spatial resolutions, typically down to 0.5–1 µm. It utilizes a to illuminate the sample, detecting of light (the Raman effect) to reveal molecular vibrations that provide a unique "fingerprint" of the material's , structure, phase, and crystallinity without requiring . The technique is grounded in the Raman effect, discovered in 1928 by Indian physicist , who observed that a very small fraction (about 0.00001%, or 1 in 10 million) of incident monochromatic light scatters inelastically, shifting in wavelength based on the sample's molecular bonds and vibrations. In a Raman microscope, this is achieved through confocal optics, high-numerical-aperture objectives (e.g., N.A. 0.75 for a ~0.44 µm spot size at 514 nm wavelength), and a spectrometer coupled with a (CCD) detector to collect and analyze the scattered light spectrum. Key advantages include its versatility for solids, liquids, gases, and aqueous solutions; insensitivity to ; and ability to probe subsurface features or stress/strain in materials (e.g., a 10 cm⁻¹ spectral shift per 1% strain in ). Raman microscopy was pioneered in the 1970s in by researchers Michel Delhaye and Paul Dhamelincourt, with the first commercial instrument, the (Molecular Optical Laser Examiner), introduced by what is now HORIBA Scientific. It has since evolved with advancements in technology (common wavelengths: 244 nm, 514 nm, 785 nm), automated mapping stages for chemical imaging (e.g., scanning 9 mm × 16 mm areas with thousands of spectra in minutes), and integration with other modalities like scanning microscopy for enhanced correlative analysis. Applications span diverse fields, including for semiconductors and polymers, pharmaceuticals for drug formulation and polymorphism, life sciences for cellular imaging, for mineral identification, forensics, and art conservation, where it enables label-free, characterization of complex, heterogeneous samples.

Fundamentals

Principles of Raman Spectroscopy

The Raman effect refers to the of monochromatic light by s, where the scattered s experience a change in energy due to interactions with molecular vibrational, rotational, and other low-frequency modes. This phenomenon, first observed in , results in two primary types of shifts: Stokes scattering, where the scattered light has lower energy (longer wavelength) than the incident light as the molecule gains energy from the photon, and anti-Stokes scattering, where the scattered light has higher energy (shorter wavelength) as energy is transferred from an already excited molecule to the photon. The energy shift, known as the Raman shift, is quantified in wavenumbers (cm⁻¹) by the equation \Delta \nu = \nu_0 - \nu_1, where \Delta \nu is the Raman shift, \nu_0 is the wavenumber of the incident light, and \nu_1 is the wavenumber of the scattered . For Stokes lines, \Delta \nu is positive, corresponding directly to the energy difference between molecular vibrational states, while anti-Stokes lines exhibit a negative shift of equal magnitude but are typically weaker at due to fewer molecules in excited vibrational states. This shift provides a direct measure of the molecular energy levels involved. Raman activity follows specific selection rules: a vibrational mode is Raman-active if it induces a change in the molecular —the ease with which the electron cloud distorts in response to an —during the vibration. This contrasts with (IR) , where activity requires a change in the molecule's . As a result, Raman and IR are complementary techniques, with Raman often detecting symmetric modes (e.g., homonuclear diatomic stretches) that are IR-inactive. The Raman spectrum acts as a molecular , with unique band positions arising from specific vibrations; for instance, the C-H modes in compounds appear around 2900 cm⁻¹, while the amide I band (primarily C=O stretch in proteins) occurs near 1650 cm⁻¹. These signatures enable precise identification of and structures without . However, the Raman effect is inherently weak, with a typical cross-section on the order of 10⁻³⁰ cm² per per , necessitating high-intensity sources to achieve detectable signals.

Historical Development

The Raman effect was discovered in 1928 by Indian physicist Chandrasekhara Venkata Raman, who observed that a small portion of light scattered by molecules undergoes a shift due to , providing a means to probe molecular vibrations. For this groundbreaking work, Raman was awarded the in 1930, marking the inception of what would become . In the decades following the discovery, advanced without lasers, relying on intense light sources like mercury arc lamps for excitation during through , with detection via photographic plates or early spectrophotometers. These early instruments cataloged molecular vibrational frequencies but suffered from weak signals and long exposure times, limiting widespread adoption. The of the continuous-wave helium-neon (He-Ne) in revolutionized the field by providing a coherent, monochromatic light source, enabling the first commercial laser-based Raman spectrometers in the mid-1960s and dramatically improving signal intensity and . Raman microscopy emerged in the 1970s through the integration of with optical , particularly with the development of confocal optics to achieve on the micrometer scale; French researchers Michel Delhaye and Paul Dhamelincourt at the built the first Raman microprobe, known as the MOLE (Molecular Optic Laser Examiner), around 1975. The first commercial Raman microscopes appeared in the 1980s from Instruments SA (now part of HORIBA), commercializing these designs for broader scientific use. Subsequent milestones included the introduction of fully confocal Raman microscopes in the 1990s, such as HORIBA's LabRAM system in 1993, which enhanced depth and enabled three-dimensional chemical . In the 2000s, tip-enhanced (TERS) variants emerged, with the first experimental demonstrations in 2000 using scanning probe tips to achieve nanoscale through plasmonic enhancement. The 2010s saw advances in stimulated () microscopy, first reported in 2008, which provided faster, background-free by amplifying Raman signals coherently for live-cell and biomedical applications. More recently, post-2020 developments have focused on accessibility, exemplified by the Open Raman Microscopy (ORM) framework introduced in a 2025 , offering modular, and software for customizable, cost-effective Raman systems.

Instrumentation and Configuration

Optical Components and Setup

A Raman microscope integrates a Raman spectrometer with an optical microscope to enable spatially resolved chemical analysis at the micrometer scale. The core hardware includes a laser source for excitation, a beam expander to adjust the beam diameter, a dichroic mirror to direct the incident light, a high-numerical-aperture (NA) objective lens for focusing, a sample stage for positioning, a notch filter to suppress Rayleigh scattering, a spectrometer for dispersing the signal, and a charge-coupled device (CCD) detector for recording the spectrum. This setup typically employs either an upright or inverted microscope configuration, with upright systems suitable for opaque or thick samples and inverted ones ideal for transparent or liquid-containing specimens. High-NA objectives, such as 100× with NA 1.4, focus the laser beam to the diffraction limit, achieving spot sizes of approximately 200–500 nm, which defines the spatial resolution for sample interrogation. The beam path in a standard Raman microscope follows a 180° backscattering to maximize signal collection efficiency. The beam passes through the beam expander to match the objective's , then reflects off the dichroic mirror toward the objective lens, which tightly focuses it onto the sample. The inelastically scattered Raman light, along with elastic , is collected back through the same objective and transmitted through the dichroic mirror, which is designed to reflect the excitation wavelength while passing longer wavelengths. A notch filter subsequently removes the intense component, allowing the weaker Raman signal to enter the spectrometer for wavelength dispersion and analysis by the detector. Sample handling in Raman microscopes relies on precision stages that enable controlled or translation for mapping across regions of interest. These stages, often motorized and computer-controlled, accommodate a variety of sample types, including solids, powders, and biological materials, with minimal preparation required. For aqueous or hydrated samples, objectives—such as water or oil types—minimize mismatches, reducing aberrations and improving depth penetration. Safety and alignment procedures are integral to Raman microscope operation due to the use of high-power lasers classified as Class 3B or 4. Enclosures, interlocks, and protective eyewear are standard to prevent exposure, while automated alignment systems use beam-steering optics and feedback mechanisms to ensure the laser focuses precisely at the sample plane, optimizing signal-to-noise ratios. Proper alignment of the spectrometer entrance slit with the focused beam is critical to avoid signal loss and maintain spectral fidelity.

Laser Sources and Detectors

Laser sources in Raman microscopy primarily consist of solid-state, , and tunable lasers, which provide the monochromatic excitation light necessary for generating Raman scattering signals. Solid-state lasers, such as frequency-doubled Nd:YAG lasers emitting at 532 nm, are commonly used for their stability and efficiency in visible-range applications. lasers, particularly those at 785 nm, offer compact, cost-effective options with low power consumption, making them suitable for routine microscopy setups. Tunable lasers, like Ti:Sapphire systems, allow flexibility across a broad spectrum (e.g., 700-1000 nm) for specialized experiments requiring variable excitation. Common wavelengths include 532 nm (green), 785 nm (near-infrared), and 1064 nm, selected based on the sample's to optimize signal intensity while minimizing interference. The cross-section scales with the of the , favoring shorter wavelengths like 532 nm for stronger signals in non-fluorescent materials, but longer wavelengths such as 785 nm or 1064 nm are preferred for biological samples to reduce autofluorescence from electronic transitions. For instance, near-infrared (e.g., 785 nm) provide a balance between efficiency and low fluorescence background, essential for imaging organic tissues without sample damage. Laser power at the sample typically ranges from 1 to 100 mW, with 5-50 mW common to achieve sufficient signal without inducing or heating effects. Detection in Raman microscopy relies on spectrometers and sensitive array detectors to disperse and capture the weak inelastically scattered light. Grating-based spectrometers in Czerny-Turner configuration, featuring an entrance slit, two mirrors, a , and an exit port, are standard for their high throughput and ability to resolve Raman shifts. These systems achieve spectral resolutions of approximately 1-5 cm⁻¹, determined by grating groove density (e.g., 1200 lines/mm) and slit width, enabling clear separation of vibrational bands. Charge-coupled devices (CCDs) serve as primary detectors due to their high (>90% in the visible range) and low noise, ideal for integrating the dispersed across multiple pixels. For low-light conditions prevalent in Raman imaging, electron-multiplying CCDs (EMCCDs) enhance through on-chip amplification, achieving sub-electron readout and detecting signals at exposure times as short as 5 ms. Array detectors like CCDs and EMCCDs enable parallel acquisition of the full , outperforming single-point detectors in speed and for applications. Scientific (sCMOS) detectors are increasingly used as alternatives, offering faster readout rates (>50 Hz) and larger for brighter signals, though with slightly higher than EMCCDs in ultra-low-light scenarios. Noise reduction is critical for reliable Raman detection, with thermoelectric (Peltier) cooling of detectors reducing dark current—thermally generated electrons that mimic signal—by 50% for every 5-7°C temperature drop. Cooled CCDs and EMCCDs, often maintained at -30 to -70°C below ambient using single- or two-stage Peltier elements, minimize dark noise to negligible levels, enhancing signal-to-noise ratios in quantitative biological imaging. This cooling also suppresses at low frequencies, allowing longer integration times without thermal saturation.

Imaging Modes

Point Illumination and Scanning

In point mapping mode, the beam is focused to a diffraction-limited spot of approximately 1 μm in diameter on the sample surface, allowing for precise localized excitation of . The illumination point is then raster-scanned across the region of interest using either a motorized translation stage or galvanometer-controlled mirrors to systematically cover the area by . This sequential acquisition approach typically requires an integration time of seconds to minutes per point to achieve sufficient , depending on the , sample properties, and desired spectral quality. The resulting dataset forms a hyperspectral datacube, where each spatial position (x, y) is associated with a full Raman spectrum across wavelengths (λ), enabling multidimensional chemical analysis. Post-acquisition, spectral unmixing algorithms are applied to decompose the datacube into constituent chemical components, facilitating the generation of false-color chemical maps that highlight spatial distributions of specific molecular species. This mode offers high signal-to-noise ratios for detecting weak Raman signals, as the entire power is concentrated on a single point, maximizing collection efficiency from that location. It is particularly well-suited for heterogeneous samples, where fine spatial variations in composition require targeted, high-fidelity measurements to resolve subtle chemical gradients. Software processing is essential for , including baseline correction to remove fluorescence-induced offsets and cosmic ray removal to eliminate spurious spikes from high-energy particles. Point illumination and scanning are commonly employed for high-resolution chemical profiling in two-dimensional surface mappings and three-dimensional volumetric reconstructions, such as delineating distributions in composite materials or tracking molecular changes within layered structures.

Wide-Field and Global Imaging

In wide-field Raman microscopy, an expanded beam illuminates the entire simultaneously, enabling parallel acquisition of Raman signals from multiple spatial points without mechanical scanning. This approach typically employs a continuous-wave focused at the back focal plane of the objective to achieve homogeneous illumination over areas up to 110 × 110 μm². Wavelength selection is facilitated by tunable filters, such as tunable filters (LCTFs), which allow continuous spectral tuning (e.g., 420–730 nm) with a bandwidth of about 10 nm (300 cm⁻¹), isolating specific Raman bands like the peak at 520.7 cm⁻¹. Detection occurs via high-sensitivity 2D array detectors, such as electron-multiplying charge-coupled devices (EMCCDs) with 512 × 512 pixels operating at up to 56 frames per second, cooled to minimize noise. Global Raman imaging extends this parallelism by incorporating spatial light modulators (SLMs) or structured illumination to enhance resolution and throughput in full-field detection setups. For instance, structured illumination Raman microscopy (SIRM) uses a (DMD) as an SLM to project interference fringes onto the sample, modulating high-frequency spatial information that is later demodulated via phase-shift algorithms for down to 80 spatially and 50 cm⁻¹ spectrally. Full-field detection relies on 2D array detectors like sCMOS cameras (e.g., ORCA-Flash 4.0), capturing across the entire illuminated area in a single exposure or a few sequential frames. This method supports excitation tailored with sparse-sampling masks in variants like spatial light-modulated stimulated (SLM-SRS). These techniques offer significant speed advantages over point illumination scanning, which requires sequential pixel-by-pixel acquisition and can take hours for large fields, by enabling image acquisition in seconds to minutes—for example, 9 seconds for a 32 × 32 μm² megapixel image in SIRM or 38 minutes for a 512 × 512 Raman using Fourier-transform approaches. However, the parallel nature reduces signal intensity per due to distributed power, leading to lower (e.g., 2.7 × 10³ W/cm² versus 10⁶ W/cm² in scanning) and potential trade-offs in or spectral fidelity, often necessitating multiple acquisitions (e.g., 9 frames) for reconstruction. A common variant, line-scanning Raman microscopy, uses slit illumination to project a line-shaped beam across the sample, combined with one-axis mechanical scanning, providing a compromise between the full parallelism of wide-field methods and the high signal-to-noise of point scanning. This setup focuses through a spectrometer slit onto a detector, achieving imaging speeds over 100 times faster than confocal scanning (e.g., 185 seconds per frame versus hours) while maintaining diffraction-limited resolution. Data handling in these modes often involves multivariate analysis for image reconstruction, such as (PCA) for and clustering of hyperspectral datasets, or multivariate curve resolution (MCR) to decompose mixed spectra into pure components and concentration maps, enhancing contrast and specificity in complex samples.

Resolution and Limitations

Spatial and Spectral Resolution

The spatial resolution of a Raman microscope is fundamentally limited by , governed by the excitation λ and the NA of the objective lens. The lateral resolution is approximately λ/(2 NA), yielding typical values of 250–500 nm for visible lasers (λ ≈ 500–800 nm) with high-NA objectives (NA ≈ 0.9–1.4). Axial resolution, which determines depth selectivity, is poorer and scales as λ/NA², typically achieving 1–2 μm under similar conditions due to the elongated focal volume along the . Spectral resolution in Raman microscopy refers to the spectrometer's ability to distinguish closely spaced Raman shifts, primarily determined by the diffraction 's groove density (lines per ) and the illuminated grating area. The minimum resolvable wavenumber shift Δν is approximated by Δν = 1/(d sin θ), where d is the grating spacing (d = 1/groove density) and θ is the diffraction angle; practical systems with 600–1800 grooves/ gratings and narrow slits achieve 1–10 cm⁻¹ . This underpins chemical resolution, enabling the identification of distinct molecular species through separation of vibrational peaks differing by as little as 10 cm⁻¹, such as the symmetric and asymmetric stretches in ions. Standard calibration of spectral resolution often employs beads or films, whose well-characterized Raman bands (e.g., at 1001 cm⁻¹) serve as traceable references for verifying instrument performance per ASTM E1843 guidelines.

Factors Affecting Resolution

Autofluorescence represents a primary limiting in Raman microscopy, arising from the excitation of endogenous in biological samples, such as aromatic amino acids like and , which produce broad bands overlapping the Raman signal. This background significantly reduces the signal-to-background (SBR), often by orders of magnitude in pigmented or organic-rich specimens. Mitigation strategies include shifting to longer excitation wavelengths, such as 1064 nm, which minimize fluorophore excitation and can improve the SBR by reducing fluorescence intensity compared to shorter wavelengths like 785 nm, thereby enhancing effective in challenging samples. Sample heterogeneity, particularly in turbid media like tissues, introduces scattering losses that degrade resolution by diffusing the incident and collected light, confining to approximately 100 μm in conventional setups. In such media, multiple events cause photons to follow random paths rather than ballistic trajectories, leading to signal and spatial blurring that limits the achievable detail. This effect is pronounced in biological tissues with high or cellular content, where Mie and dominate, further reducing contrast and effective resolution. Instrumental noise sources, including , read noise, and thermal effects, impose operational limits on by introducing variability in the detected Raman signal. , governed by statistics, scales with the square root of count and becomes prominent in low-signal regimes typical of . Read noise originates from detector electronics during signal readout and remains constant regardless of exposure, while thermal effects generate dark current through electron-hole pairs in the detector, exacerbating noise at elevated temperatures. These noises collectively lower the , effectively broadening spectral features and diminishing spatial fidelity during imaging. Alignment issues, such as beam walk-off or focus drift during extended scans, arise from mechanical vibrations, thermal expansions, or optical misalignments, causing deviations in the focal plane and resultant resolution loss. In long-duration acquisitions, even minor drifts can shift the excitation spot by several micrometers, leading to inconsistent sampling and artifacts in reconstructed images. Maintaining precise through active stabilization or periodic recalibration is essential to preserve resolution over time. Overall, these factors can degrade by up to twofold in biological samples compared to ideal conditions, primarily due to mismatch between the sample (typically n ≈ 1.4–1.5) and the objective's immersion medium (e.g., air n = 1 or water n = 1.33), inducing spherical aberrations that elongate the axial . This mismatch distorts the focal volume, reducing both lateral and depth in practice, though it aligns with theoretical spatial limits under .

Enhancement Techniques

Confocal and Tip-Enhanced Methods

Confocal Raman microscopy employs a pinhole , typically sized between 10 and 100 μm, positioned in the to reject out-of-focus light, thereby enhancing the spatial selectivity of the Raman signal collection. This configuration significantly improves axial to approximately 0.5–1 μm, compared to the diffraction-limited baseline of several micrometers in non-confocal setups, enabling optical sectioning for three-dimensional imaging of samples. The pinhole acts as a , confining the detection volume and allowing for depth-resolved analysis without physical slicing, which is particularly useful for layered or heterogeneous materials. Tip-enhanced Raman scattering (TERS) overcomes the diffraction limit of conventional Raman microscopy by using plasmonic tips, such as those coated with silver (Ag) or () nanoparticles, to localize the and amplify the Raman signal by factors of 10 to 100 times. This near-field enhancement achieves spatial resolutions below 10 nm, enabling nanoscale chemical mapping of surfaces and interfaces. TERS configurations are broadly classified into apertureless and apertured types; apertureless setups rely on scattering from the tip apex in (AFM) or (STM) modes, while apertured variants integrate a subwavelength in the tip, often combined with for precise control. Excitation schemes in TERS include normal (top or bottom) illumination, where the is directed along the tip axis, and side illumination, which focuses the beam perpendicular to the tip-sample gap to minimize background scattering. Key challenges in TERS include maintaining tip during scanning to avoid signal fluctuations and achieving reproducible enhancement across multiple probes due to variations in tip geometry and plasmonic properties. The enhancement factor G arises primarily from the electromagnetic mechanism and is given by G = \left| \frac{E_\text{local}}{E_\text{inc}} \right|^4, where E_\text{local} is the local at the tip apex and E_\text{inc} is the incident , reflecting the fourth-power dependence of Raman on the field strength. Recent developments in the 2020s have focused on chemically modified tips and to improve sensitivity, selectivity, and for applications in biomaterials and nanoscale .

Stimulated Raman Scattering

Stimulated (SRS) microscopy employs a pump-probe configuration utilizing two synchronized beams: a beam at \omega_p and a Stokes beam at \omega_s, where \omega_p > \omega_s. The difference \Delta \omega = \omega_p - \omega_s is tuned to match the vibrational of molecular bonds in the sample, inducing coherent of the Stokes beam (stimulated Raman gain, SRG) or depletion of the beam (stimulated Raman loss, SRL). This coherent process enhances the inherently weak Raman signal by orders of magnitude compared to spontaneous , enabling faster imaging without exogenous labels. The core of SRS signal generation is described by the for the change along the propagation direction z: \frac{dI_s}{dz} = \sigma I_p I_s where I_s and I_p are the of the Stokes and beams, respectively, and \sigma is the Raman cross-section, which depends on the molecular species and \Delta \omega. In practice, the small fractional change in beam (\sim 10^{-4} to $10^{-5}) is detected using techniques, such as of one beam at a kilohertz rate, followed by lock-in amplification to extract the SRS contrast against background noise. This setup allows quantitative, linear measurement of molecular concentrations, as the signal is directly proportional to the of vibrators. A key variant of coherent Raman techniques is coherent anti-Stokes Raman scattering (CARS), which generates a blueshifted anti-Stokes signal but suffers from a non-resonant background that distorts spectral lineshapes and limits quantitative accuracy. In contrast, SRS provides a purely resonant, linear response without this background, making it preferable for precise chemical identification and quantification in complex samples. SRS achieves imaging speeds 10⁴ to 10⁶ times faster than spontaneous Raman microscopy due to the coherent signal amplification, facilitating video-rate acquisition (up to 30 frames per second) with minimal photodamage. Implementations often leverage lasers for SRS, where chirped pulses enable over several hundred wavenumbers (e.g., 2800–3000 cm⁻¹ for C-H stretches) in a single scan, achieving resolutions of ~10 cm⁻¹. Spatial resolution in SRS microscopy is diffraction-limited at approximately 300 nm laterally and 1 µm axially under high-numerical-aperture objectives, sufficient for subcellular . These systems support point-scanning or wide-field configurations, with the coherent reducing acquisition times to microseconds per . Recent advances in 2025 have expanded applications in , particularly for live-cell , with innovations like photothermal detection-based microscopes achieving sub-300 nm resolution and 10-fold sensitivity gains for tracking metabolic dynamics and drug interactions in . implementations have further improved , enabling prolonged observation of cellular processes such as without compromising viability. These developments, including novel vibrational probes, enhance 's role in discovery and multiplexed of endogenous biomolecules.

Applications

Biological and Biomedical Uses

Raman microscopy enables label-free of biomolecules in biological systems, providing insights into molecular without the need for exogenous labels that can alter cellular processes. In live-cell applications, stimulated Raman scattering (SRS) variants allow high-speed visualization of lipid distribution and protein secondary structures within organelles, facilitating the study of dynamic metabolic processes. For instance, SRS microscopy has been used to track lipid droplet dynamics in adipocytes and neurons, revealing alterations in associated with cellular stress. Additionally, quantitative SRS distinguishes protein secondary structures, such as alpha-helices and beta-sheets, in phase-separated condensates, offering a window into protein misfolding mechanisms relevant to cellular function. In tissue analysis, Raman microscopy supports histology by mapping spectral signatures of key biomolecules, aiding in diagnostics. For cancer detection, differences in Raman peaks, such as the 1080 cm⁻¹ band attributed to nucleic acids and the 1440 cm⁻¹ band for , distinguish malignant from healthy tissues with high specificity in models of and . This approach enables rapid, non-destructive assessment of tumor margins and heterogeneity, improving pathological evaluation over traditional staining methods. Raman microscopy plays a crucial role in monitoring systems, particularly the uptake and release of therapeutics from nanoparticles in cellular environments. Label-free imaging tracks the intracellular distribution of drug-loaded nanocarriers, quantifying uptake and endosomal escape in without . For example, studies have visualized the release of chemotherapeutic agents from polymeric nanoparticles in cancer cells, correlating changes with therapeutic . The potential of is realized through endoscopic probes that provide real-time molecular guidance during . Fiber-optic integrated with endoscopes deliver multiplexed spectral data for characterization, enabling intraoperative delineation of tumor boundaries in procedures like . These systems achieve detection sensitivities comparable to . In neurodegeneration research, 2025 advancements in have highlighted its utility in modeling and dysregulation. For instance, SRS has imaged protein aggregates, such as β-sheet formation in FUS condensates and Htt models, quantifying alterations relevant to neurodegenerative diseases like Alzheimer's. Unlike , Raman techniques avoid , allowing prolonged imaging of live tissues without signal degradation or sample damage. Raman microscopy extends to microbiome analysis and viral detection, offering rapid, non-destructive identification of microbial communities and pathogens. In microbiome studies, it differentiates bacterial species based on vibrational fingerprints, such as cell wall compositions in and genera, enabling label-free profiling of gut or soil samples. For viral detection, (SERS) variants identify respiratory viruses, including , in biological and clinical samples with high accuracy (>90%), leveraging viral protein signatures.

Materials and Chemical Analysis

Raman microscopy plays a crucial role in materials and chemical analysis by providing non-destructive, spatially resolved chemical and of inorganic, polymeric, and other non-biological materials. This technique leverages the molecular vibrational fingerprints obtained from to map composition, structure, and properties at the microscale, enabling detailed insights into material heterogeneity without sample preparation that could alter the specimen. In analysis, Raman microscopy is widely used for mapping in wafers, where shifts in peaks, such as the silicon transverse optical mode around 520 cm⁻¹, quantify local stress distributions induced by processing or device fabrication. For instance, this method has revealed gradients in bent gallium phosphide nanowires, correlating peak shifts with tensile or compressive forces to assess mechanical integrity. Additionally, phase identification in alloys benefits from Raman's ability to distinguish crystalline phases through characteristic vibrations; in aluminum-silicon coated steels, cross-sectional mapping identifies phases like Al₃FeSi post-heat treatment, aiding in . Polymer science applications include assessing crystallinity and chain orientation in plastics, where polarized Raman spectroscopy measures depolarization ratios of specific bands to determine the degree of order. In poly(p-phenylene terephthalamide) fibers, the I band at ~1650 cm⁻¹ intensity variations quantify crystallinity levels, linking them to mechanical properties like tensile strength. Degradation studies track changes in peak positions and widths, such as carbonyl stretches in oxidized , to monitor environmental aging without destroying the sample. In the , Raman microscopy excels at polymorph detection in drug formulations, identifying versus hydrate forms via distinct Raman peaks; for example, polymorphs are differentiated by linearly and circularly polarized spectra, ensuring consistent bioavailability. This extends to tablet uniformity mapping, where scans active pharmaceutical ingredients across cross-sections to detect distribution variations at the micrometer scale. Forensic applications utilize Raman microscopy for trace evidence analysis, such as identifying inks and through their unique spectral signatures, and mapping explosive residues on surfaces. Inks are distinguished by colorant vibrations, as seen in traces on fabrics, while fiber composition is confirmed via polymer bands; explosive residues, like those from , are mapped using peaks around 1040 cm⁻¹ to reconstruct handling non-destructively. Environmental monitoring employs Raman microscopy for microplastics identification in sediments, where spectral libraries match particle signatures to polymers like (peaks at 1060 and 2880 cm⁻¹). This enables rapid classification and sizing, crucial for assessing impacts. Catalyst surface further applies the technique to visualize active sites and intermediates, with tip-enhanced variants probing adsorbates on metal nanoparticles to optimize performance. Quantitative analysis in Raman microscopy relies on calibration curves derived from peak intensities versus known concentrations, allowing concentration mapping of analytes in heterogeneous materials. For microplastics in water, internal standard peaks normalize signals to achieve detection limits below 1 mg/L, facilitating accurate distribution maps in complex matrices. Chemical fingerprinting underpins these identifications, providing a basis for multivariate analysis of spectral datasets.

Advanced and Correlative Techniques

Multimodal and Correlative Imaging

Multimodal imaging in Raman microscopy involves integrating with complementary techniques to provide a more comprehensive of samples, combining data from Raman with structural, functional, or elemental information from other modalities. This approach addresses limitations of standalone Raman, such as its moderate , by leveraging the strengths of paired methods like fluorescence microscopy for high-contrast visualization or electron microscopy for nanoscale topography. Correlative imaging specifically refers to the overlay or co-registration of datasets from these modalities, often achieved through sequential or simultaneous acquisition on the same sample region, enabling spatially resolved correlations between molecular vibrations and other properties. In biological and biomedical applications, Raman is frequently combined with microscopy to distinguish endogenous fluorophores from Raman signals, improving specificity in tissue diagnostics. For instance, near-infrared autofluorescence-Raman multimodal systems have been used for real-time gastric cancer detection, achieving high by correlating fluorescence intensity with Raman spectral fingerprints of cellular components. Similarly, Raman-optical coherence (OCT) hybrids enhance discrimination of colonic , with reported sensitivities up to 94%, as OCT provides depth-resolved structural images that guide Raman's chemical mapping. These combinations reduce false positives in surgical margins assessment, such as for , by integrating rapid morphological data with label-free molecular insights. For materials characterization, correlative Raman imaging with scanning electron microscopy (SEM) or (TEM) allows simultaneous chemical and morphological analysis at the micro- to nanoscale. In energy materials like carbon nanotube composites, SEM-Raman reveals phase distributions and defects by overlaying Raman's vibrational spectra with SEM's topographic contrast, aiding in performance optimization. Tip-enhanced Raman scattering (TERS), often paired with (AFM), pushes resolution below 1 nm via plasmonic enhancement, correlating nanoscale chemical heterogeneity with surface topography in polymers and . Such multimodal setups are particularly valuable for non-destructive analysis of heterogeneous samples, like bacterial biofilms, where TEM complements Raman to map elemental and molecular distributions. Advanced correlative workflows extend to indirect multimodal bioimaging, such as combining MRI or with Raman on tissue sections. This has been applied to models, where tracks tumor heterogeneity in living subjects, followed by Raman imaging of cleared brain slices to reveal single-cell chemical signatures, enhancing understanding of disease progression. Advantages include bridging scales from organismal to molecular levels, though challenges like data co-registration persist, often addressed via fiducial markers or software alignment. Overall, these techniques have revolutionized fields like and by providing holistic, high-fidelity datasets.

Recent Advances and Future Directions

Recent advances in Raman microscopy have been driven by integrations of (AI) and (ML), particularly for spectral processing and analysis. Deep learning models, such as convolutional neural networks (CNNs), have enabled effective spectral denoising, improving signal-to-noise ratios (SNRs) by up to 57-fold for chemical samples and 15-fold for biological samples like C. elegans in stimulated Raman scattering (SRS) imaging. Automated classification pipelines, including support vector machines (SVMs) and ensemble methods, facilitate high-throughput single-cell identification in hyperspectral datasets, enhancing diagnostic accuracy for conditions like brain tumors. CNN-based unmixing techniques, such as variants, reconstruct chemical maps from low-power, high-speed images, achieving SNR levels comparable to 100-fold averaging and supporting label-free biomolecular analysis at cellular resolution. Quantum-enhanced approaches have addressed noise limitations in Raman microscopy through the use of squeezed light, reducing photon-number fluctuations and enabling faster, lower-damage imaging. A 2025 platform employing amplitude-squeezed light (5.2 dB noise reduction) in SRS microscopy covers broad spectral ranges (1000–3100 cm⁻¹), including fingerprint and CH-stretch regions, with a 51% SNR improvement while minimizing photodamage in biological tissues like pork muscle. This quantum enhancement suppresses noise by an average of 3.6 dB, offering greater chemical specificity for metabolite visualization. Super-resolution techniques have pushed spectral resolution boundaries in X-ray Raman microscopy. In a 2025 study, stochastic stimulated X-ray Raman scattering (s-SXRS) using self-amplified spontaneous emission (SASE) X-ray free-electron laser pulses achieved 0.1 spectral resolution and 40 fs temporal resolution, surpassing the spectrometer's 0.18 instrumental broadening and compressing the 7.5 SASE linewidth by 80-fold via covariance analysis. This enabled distinction of sub-eV-separated valence-excited states (e.g., 0.31 splitting in neon-like ions), providing femtosecond snapshots of site-specific dynamics in complex systems. Open-source initiatives have democratized access to Raman microscopy through modular designs. The Open Raman Microscopy (ORM) framework, introduced in 2025, provides a cost-effective, customizable with interchangeable components and integrated software for label-free chemical . This setup supports high-throughput configurations for diverse applications, reducing barriers for researchers in resource-limited settings. Looking ahead, developments in portable in vivo systems promise broader clinical applicability. Handheld confocal Raman probes and fiber-based SRS setups enable non-invasive, real-time tissue assessment, such as in dermatological and endoscopic procedures, with ultrawideband capabilities for in vivo measurements. However, clinical translation faces challenges including standardization, precise alignment, validation against gold standards, and integration into workflows to ensure reliability across diverse patient populations.

References

  1. [1]
    What is Raman Spectroscopy? - HORIBA
    Raman Spectroscopy is a non-destructive chemical analysis technique based on light scattering, providing detailed information about chemical structure.
  2. [2]
    None
    ### Summary of Raman Microscopy
  3. [3]
    Renishaw: Raman spectroscopy
    ### Summary of Raman Spectroscopy and Its Use in Microscopy
  4. [4]
    5: Raman Spectroscopy - Chemistry LibreTexts
    Mar 16, 2023 · Anti-Stokes lines are scattered photons that are increased in energy relative to the incident photons that interacted with the molecule. The ...Learning Objectives · Consider the molecular... · What effect would raising the...
  5. [5]
    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.
  6. [6]
    What is Raman Spectroscopy? - StellarNet, Inc.
    Feb 26, 2017 · This is calculated by the following equation: where Δν~ is the shift in wavenumbers, λ0, is the laser wavelength, and λ1 is the scattered ...Missing: ν₀ - ν₁
  7. [7]
  8. [8]
    1.13: Selection Rules for IR and Raman Spectroscopy
    Feb 23, 2025 · Selection rules such as these are used to tell us whether such transitions are allowed, and therefore observed, or whether they are forbidden.
  9. [9]
    [PDF] Comparison between IR and Raman
    IR Selection Rule: For any IR absorption, the vibrational excitation must change the dipole moment of the molecule in order for radiation absorption to occur.
  10. [10]
    Interpretation of Raman Spectrum of Proteins | Spectroscopy Online
    Feb 1, 2022 · The methylene group often has symmetric and asymmetric stretches respectively near 2850–2900 cm -1, whereas a methyl group has symmetric and ...
  11. [11]
    Identification of Individual Components from the Manufacturing ...
    Jul 30, 2016 · The drug substance contains proteins which gave prominent Raman peaks for C-H stretch and amide I at 2900 and 1650 cm-1, respectively. This ...Missing: examples | Show results with:examples
  12. [12]
    An Einstein-coefficient-like equation for Raman cross sections - PMC
    Nov 15, 2023 · The values of σRaman (10−30 to 10−28 cm2 for small chemical bonds) turn out to be many orders of magnitude smaller in comparison to other ...
  13. [13]
    Sir Venkata Raman – Facts - NobelPrize.org
    In 1928 Venkata Raman discovered that a small portion of the scattered light acquires other wavelengths than that of the original light.
  14. [14]
    Discover 50 years of Raman innovation by HORIBA
    Raman spectroscopy was named after Sir Chandrasekhara Venkata Raman (7 November 1888 – 21 November 1970), an Indian physicist born in the former Madras ...
  15. [15]
    A History of Raman Spectroscopy - BCC Research Blog
    Jan 29, 2019 · In the years following its discovery, Raman spectroscopy was used to provide the first catalog of molecular vibrational frequencies.
  16. [16]
    A History of the Laser: 1960 - 2019 | Features - Photonics Spectra
    December 1960: Ali Javan, William Bennett Jr. and Donald Herriott of Bell Labs develop the helium-neon (HeNe) laser, the first to generate a continuous beam of ...
  17. [17]
    Principles of Raman spectroscopy (4) History of ... - JASCO Global
    Dec 9, 2021 · In the 1960s, laser Raman spectrophotometers using a 694.3 nm ruby ​​laser and a 632.8 nm He-Ne laser appeared, although the output power ...
  18. [18]
    Raman Micro Imaging - What Was a Concept in 1975 Is Now a Reality
    Michel Delhaye at the Technical University of Lille in northern France recognized the scientific potential of a Raman microscope (1). There apparently had been ...
  19. [19]
    [PDF] My career path in the field of Raman spectroscopy and microscopy
    Mar 14, 2022 · Instruments SA/Jobin Yvon who had recently introduced the Raman microscope designed in Lille, France by Professors. Michel Delhaye, Paul ...Missing: history | Show results with:history
  20. [20]
    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).
  21. [21]
    Principle of Stimulated Raman Scattering Microscopy - NIH
    Jun 30, 2025 · Stimulated Raman scattering (SRS) microscopy, on the other hand, was first reported in 200813–15. SRS microscopy leverages the stimulated Raman ...
  22. [22]
    Open Raman Microscopy (ORM): A Modular Hardware and Software ...
    Oct 26, 2025 · Posted October 26, 2025. Download PDF. Print/Save ... Follow this preprint. X. You can now receive automatic notifications when a preprint ...
  23. [23]
    Instrument Presentation - HORIBA
    A Raman microscope combines a Raman spectrometer with a standard optical microscope. The excitation laser beam is focused through the microscope to create a ...
  24. [24]
    Raman Spectroscopy Setup: What are the Key Components and ...
    The spectral range of the Raman signal is shaped by optical components such as dichroic filters and long-pass filters. Portable Raman Spectrometer. Beam ...
  25. [25]
    Basic Aspects of Experimental Design in Raman Microscopy
    Nov 1, 2016 · This article discusses basic practical aspects of analytical instrument performance and experimental design that should be taken into consideration
  26. [26]
    Guide to Raman Microscopy | Bruker
    We briefly explain the fundamentals of Raman microscopy and take a closer look on why things like spectral resolution and confocality are so important.
  27. [27]
    Lasers for Raman Spectroscopy - RP Photonics
    Short wavelengths are very much favored by the Raman scattering efficiency, which typically scales with the fourth power of the optical frequency. Lasers ...
  28. [28]
    How to choose lasers for Raman Spectroscopy - HÜBNER Photonics
    The most commonly used wavelength in Raman spectroscopy is 785 nm. It offers the best balance between scattering efficiency, influence of fluorescence, detector ...
  29. [29]
    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.
  30. [30]
    Design Simulation of Czerny–Turner Configuration-Based Raman ...
    We report the design simulation of the Raman spectrometer using Zemax optical system design software. The design is based on the Czerny–Turner configuration.
  31. [31]
    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.
  32. [32]
    CMOS and CCD Detection in Raman Spectroscopy: A Comparison ...
    Jan 28, 2025 · We evaluate sCMOS sensors for multiplex (50 ps) CARS and spontaneous Raman spectroscopy, making a direct comparison with a state-of-the-art EM-CCD detector.
  33. [33]
  34. [34]
    Noninvasive molecular imaging of small living subjects ... - PNAS
    Raman images were obtained by using a Raman point-mapping method. The computer-controlled x–y translation stage performed a raster scan of a region of ...
  35. [35]
    [PDF] Improving Acquisition Speed via Line Scan Raman Microscopy for ...
    Currently, the point mapping scan of 30x30 points by our Raman spectroscopy system takes approximately 30 mins. In case of performing RS on biological samples, ...
  36. [36]
    Hyperspectral Imaging with Stimulated Raman Scattering by ...
    In recent years, stimulated Raman scattering (SRS) microscopy has emerged as an alternative to CARS microscopy that is free from the aforementioned issues. (15- ...
  37. [37]
    Hyperspectral unmixing for Raman spectroscopy via physics ... - PNAS
    Hyperspectral unmixing methods are essential to exploit the capabilities of Raman spectroscopy for nondestructive, unbiased chemical characterization in a ...
  38. [38]
    raman knowledge base
    The confocal Raman imaging technique combines Raman spectroscopy with a confocal microscope by acquiring the information of a complete Raman spectrum at every ...
  39. [39]
    Fast and quantitative 2D and 3D orientation mapping using Raman ...
    Dec 5, 2019 · Confocal Raman microscopy is a powerful non-destructive technique for chemical mapping of organic and inorganic materials. Here we demonstrate ...Results · Raman Beam Delivery System... · Optical Path And Components...
  40. [40]
    Complex evaluation of Raman spectra using morphological filtering ...
    Sep 6, 2022 · We discuss a full-scale treatment of real-life Raman spectra with pronounced artifacts via mathematical morphology: Baseline correction, peak recognition, and ...
  41. [41]
    Open-sourced Raman spectroscopy data processing package ... - NIH
    Feb 21, 2023 · Cosmic Ray Removal. Cosmic rays randomly hit spectrometer camera pixels during the acquisition of Raman signals, resulting in the appearance of ...
  42. [42]
    Recording Raman spectral images and profiles - HORIBA
    Raman profiles and images can be collected in one, two and three dimensions, including: 1D: line profile, depth profile, time profile, temperature profile; 2D: ...
  43. [43]
    Spatial Resolution in Raman Spectroscopy - Edinburgh Instruments
    Jul 8, 2021 · Shorter wavelength lasers can be used to achieve higher spatial resolution. Higher NA objectives, especially immersion types, improve resolution.Key Points · Lateral Resolution · Rm5Missing: visible | Show results with:visible
  44. [44]
    [PDF] Spectral Resolution - HORIBA
    Typical gratings used in Raman spectrometers vary from about 300 gr/mm (low resolution) up to 1800 gr/mm (high resolution); more specialized gratings, such as ...
  45. [45]
    Peak Shape and Closely Spaced Peak Convolution in Raman Spectra
    Sep 1, 2021 · Peak shapes in Raman spectra are often Gaussian or Lorentzian. Closely spaced peaks can affect the convolved peak's appearance. Solids tend to ...<|separator|>
  46. [46]
    Super-resolution stimulated X-ray Raman spectroscopy - Nature
    Jul 16, 2025 · Here we demonstrate SXRS with spectroscopic precision, that is, detection of valence-excited states in neon with a near Fourier-limited joint ...
  47. [47]
    A New Fluorescence Suppression Method for Raman ... - Nature
    Oct 31, 2019 · Here we report a new approach for near complete elimination of laser-induced background fluorescence in highly pigmented biological specimens.
  48. [48]
    1064-nm Raman: The Right Choice for Biological Samples?
    Jun 1, 2017 · 1064-nm Raman vastly outperforms 785-nm Raman for background fluorescence reduction, even as compared to lengthy sample pretreatment at 785 nm via ...Missing: microscopy mitigation
  49. [49]
    Sensitivity of Transmission Raman Spectroscopy Signals to ... - Nature
    May 30, 2018 · The highly scattering nature of biological tissues makes light propagate in a random walk style, rather than directly (ballistic regime). This ...
  50. [50]
    Tissue imaging depth limit of stimulated Raman scattering microscopy
    Jan 13, 2020 · We report that for murine brain samples thinner than 2 mm transmissive imaging provides better results, while samples 2 mm and thicker are best imaged with epi ...
  51. [51]
    Understanding Raman Spectrometer Parameters
    Accurate measurement and very fine control of laser power at the tenths of milliwatts level is desirable. Some of the latest Raman systems on the market measure ...
  52. [52]
    Noise Sources and Requirements for Confocal Raman ... - MDPI
    Sensitive detectors with low readout noise and dark current shot noise are preferred for Raman applications. While the first noise source is independent of the ...
  53. [53]
    Depth Resolution of the Raman Microscope: Optical Limitations and ...
    Mar 1, 2010 · Raman microscopy has been attractive because of its ability to characterize materials on a spatial scale commensurate with optical ...Background · Results: Silicon · Results: Polystyrene Spheres
  54. [54]
    Correcting Focus Drift in Live-Cell Microscopy
    Defeating the problem of focus drift is a key consideration that must be addressed during configuration of an experimental system designed for time-lapse ...
  55. [55]
    Depth-resolved confocal micro-Raman spectroscopy for ...
    Nov 26, 2013 · Axial profile of the laser reflection from a Si sample reveals the axial resolution of our confocal micro-Raman system to be ∼0.64 μm. FIG ...
  56. [56]
    The meaning of confocal Raman microscopy - HORIBA
    For a true confocal design, typical spatial resolution is in the order of 0.5-1 µm. For a true confocal design, the limits of spatial resolution are defined ...Missing: axial | Show results with:axial
  57. [57]
    Confocal Raman Microscopy with Adaptive Optics | ACS Photonics
    The main advantage of confocal microscopes over their conventional counterparts is their ability to optically "section" thick specimens; the thin image slices ...
  58. [58]
    Tip-Enhanced Raman Spectroscopy | Analytical Chemistry
    Aug 29, 2016 · Thus, the TERS signals mostly come from the range of the gap between the tip and substrate, thus realizing a high spatial resolution of 1–10 nm.
  59. [59]
    Understanding the Role of Different Substrate Geometries for ...
    Feb 2, 2021 · Tip-enhanced Raman spectroscopy (TERS) is a powerful technique that combines Raman spectroscopy with scanning probe microscopy (SPM), i.e., ...
  60. [60]
    Investigations of Shape, Material and Excitation Wavelength Effects ...
    There are basically two kinds of TERS, which are generally accepted by the TERS community: The aperture type—using a fiber whose hollow core acts as an aperture ...
  61. [61]
    Schematic illustration of different TERS setups. (a) Bottom...
    There are three TERS configurations based on the direction of laser illumination at the sample surface, as schematically illustrated in Fig. 1.<|separator|>
  62. [62]
    Challenges and Opportunities of Tip-Enhanced Raman ...
    Sep 28, 2021 · The key challenges for TERS in liquid with the bottom setup are the production of robust tips, (51,123) the simultaneous measurements of CV and ...Introduction · Experimental Setups · The Main Experiments and... · Discussion
  63. [63]
    Tip-enhanced Raman Spectroscopy: Tip-Related Issues - PubMed
    However, TERS suffers and will continue to suffer from the limited availability of TERS tips with a high enhancement, good stability, and high reproducibility.
  64. [64]
    In vivo spectroscopy: optical fiber probes for clinical applications
    This review provides a general overview of different fiber-optic probes for in vivo measurements with emphasis on Raman spectroscopy for biomedical application.
  65. [65]
    Stimulated Raman Scattering Microscopy: A Review - MDPI
    Other advantages of SRS microscopy are high sensitivity, spatial and spectral resolution, and 3D sectioning capability.Stimulated Raman Scattering... · 2.2. Spectral Bandwidth And... · 2.5. Quantum Enhancement
  66. [66]
    Fast Vibrational Imaging of Single Cells and Tissues by Stimulated ...
    May 28, 2014 · For single-frequency imaging, an SRS microscope offers a speed that is ∼1000 times faster than a line-scan Raman microscope and 10 000 times ...Abstract · Figure 4 · 4 Quantitative Image...
  67. [67]
    Broadband hyperspectral stimulated Raman scattering microscopy ...
    Hyperspectral stimulated Raman scattering (hsSRS) microscopy has recently emerged as a powerful non-destructive technique for the label-free chemical ...
  68. [68]
    A Breakthrough SRS, Stimulated Raman microscope based on ...
    Oct 17, 2025 · The stRAMos microscope introduces a 10x sensitivity improvement over conventional SRS techniques, achieving sub-300 nm spatial resolution and ...
  69. [69]
    Fiber laser based stimulated Raman photothermal microscopy ...
    Sep 29, 2025 · The ability to perform live cell imaging while maintaining cell viability underscores the advantages of SRP over traditional methods, ...
  70. [70]
  71. [71]
    Stimulated Raman Scattering Imaging Sheds New Light on Lipid ...
    Mar 10, 2023 · This article overviews and discusses the latest applications of SRS microscopy as an emerging platform to dissect LD biology in health and disease.Figure 1 · Srs Microscopy: Technical... · Figure 2
  72. [72]
    In situ secondary structure imaging of protein phase separation and ...
    Sep 29, 2025 · Here we develop in situ quantitative imaging of secondary structure in protein condensates by stimulated Raman scattering (SRS) microscopy.
  73. [73]
    Advances in stimulated Raman scattering imaging for tissues and ...
    Stimulated Raman scattering (SRS) microscopy has emerged in the last decade as a powerful optical imaging technology with high chemical selectivity, speed, and ...Missing: history | Show results with:history
  74. [74]
    Applications of Raman spectroscopy in cancer diagnosis - PMC
    A new method using Raman spectroscopy for in vivo targeted brain cancer tissue biopsy. ... Histochemical analysis of biological tissues using Raman spectroscopy.
  75. [75]
    Raman spectroscopy for medical diagnostics — From in-vitro ...
    Jul 15, 2015 · This paper reviews the main Raman spectroscopy techniques used in medical diagnostics and provides an overview of various applications.
  76. [76]
    Strategies and Progress of Raman Technologies for Cellular Uptake ...
    Nov 20, 2023 · Nanoparticle (NP)-based drug delivery systems have the potential to significantly enhance the pharmacological and therapeutic properties of ...
  77. [77]
    Temporal imaging of drug dynamics in live cells using stimulated ...
    Aug 9, 2022 · Stimulated Raman scattering (SRS) microscopy is a powerful technique for visualising the cellular uptake and distribution of drugs and small molecules in live ...
  78. [78]
    Establishment of a clinical workflow for in vivo Raman spectroscopy ...
    Jul 7, 2025 · To overcome these issues, we mounted the Raman probe onto an endoscopic white-light camera, allowing continuous visual guidance; once the probe ...
  79. [79]
    Advances in real‐time fiber‐optic Raman spectroscopy for early ...
    Sep 25, 2020 · Huang et al [65] developed a multimodal Raman endoscope that provides intraoperative surgical guidance via wide-field (WLR/AFI/NBI) imaging ...2 Modern Raman... · 3 Fiber-Optic Raman Probe... · 7 Tissue Raman Analysis For...
  80. [80]
    Deuterium-oxide-assisted stimulated Raman scattering microscopy
    Oct 3, 2025 · In terms of application scenarios, D2O D 2 O -SRS imaging has shown significant promise in cancer biology and neurodegenerative diseases, ...
  81. [81]
    Raman spectroscopy in cell biology and microbiology - Pezzotti - 2021
    Aug 22, 2021 · This review covers miscellaneous examples of Raman analyses and imaging of eukaryotic cells, bacteria, and viruses.
  82. [82]
    Unveiling Microbial Diversity: Raman Spectroscopy's Discrimination ...
    Sep 18, 2024 · Raman spectroscopy coupled with chemometrics is a robust and helpful method for differentiating Clostridium species from Bacillus, Clostridioides, and ...
  83. [83]
    Detection of several respiratory viruses with Surface-Enhanced ...
    The study uses Surface Enhanced Raman Spectroscopy (SERS) and Artificial Intelligence (AI) to detect and differentiate three respiratory viruses.Missing: microscopy microbiome
  84. [84]
    A rapid and label-free platform for virus capture and identification ...
    Dec 27, 2019 · A portable platform that captures viruses by their size, coupled to Raman spectroscopy, resulted in successful virus identification with 90% ...
  85. [85]
    Advantages and developments of Raman spectroscopy for ... - Nature
    Oct 9, 2023 · In this review, we outline the fields of application of Raman spectroscopy and microscopy in various electroceramic systems, defining current key bottlenecks.
  86. [86]
    Investigation of strain induced effects in silicon wafers due to ...
    Sep 6, 2002 · Investigation of strain induced effects in silicon wafers due to proximity rapid thermal processing using micro-Raman spectroscopy and ...Missing: microscopy mapping
  87. [87]
    Strain Mapping and Raman Spectroscopy of Bent GaP and GaAs ...
    Mar 15, 2018 · The system involves randomly oriented SiNWs grown using established procedures on silicon wafers, and then transferred and organized into ...
  88. [88]
    Raman Spectroscopic Analysis of the Reaction between Al-Si ...
    Jul 17, 2023 · Cross-sectional Raman microscopic mapping was used to identify intermetallic phases present in Al-Si-coated 22MnB5 after heating at temperature ...
  89. [89]
    Quantifying Crystallinity in Poly(p-phenylene terephthalamide) by ...
    Jul 16, 2024 · Here we report a new method for quantifying the crystallinity in PPTA by Raman spectroscopy. The Amide I band at ∼1650 cm –1 in the Raman spectrum of PPTA was ...
  90. [90]
    Sofosbuvir Polymorphs Distinguished by Linearly and Circularly ...
    Nov 21, 2024 · This study demonstrates the potential of polarized Raman microscopy for the precise characterization of pharmaceutical polymorphs, providing a ...
  91. [91]
    Fast Compressive Raman Imaging of Polymorph Molecules and ...
    Nov 23, 2022 · The NIR compressive Raman system is successfully used to detect and image active pharmaceutical ingredients exhibiting polymorphism within compact ...Missing: microscopy detection
  92. [92]
    Raman Technique Detects Crime Scene Clues From Traces ... - C&EN
    Aug 19, 2013 · A new report shows that Raman spectroscopy, in contrast, can distinguish the brand and colors of lipstick traces in realistic crime-scene evidence.Missing: microscopy | Show results with:microscopy
  93. [93]
    Fast Detection and Classification of Microplastics by a Wide-Field ...
    Apr 29, 2025 · Here we introduce a wide-field hyperspectral Fourier transform Raman microscope for the rapid detection and identification of MPs.Introduction · Materials and Methods · Results and Discussion · References
  94. [94]
    Surface- and Tip-Enhanced Raman Spectroscopy in Catalysis
    Apr 14, 2016 · Surface- and tip-enhanced Raman spectroscopy (SERS and TERS) techniques exhibit highly localized chemical sensitivity, making them ideal for studying chemical ...Author Information · Biographies · Acknowledgment · References
  95. [95]
    Quantitative Raman analysis of microplastics in water using peak ...
    Oct 9, 2024 · This study presents a novel analytical method for the quantitative and qualitative analysis of microplastics (MPs) in deionized (DI) water using Raman ...
  96. [96]
    Surface Enhanced Raman Spectroscopy for Quantitative Analysis
    Feb 11, 2020 · The goal of calibration models is to predict the analyte concentration in an unknown (chemical) sample from instrument responses. The results ...
  97. [97]
    Looking for a perfect match: multimodal combinations of Raman ...
    Aug 12, 2021 · Here, we want to give an overview on the development of multimodal systems that use RS in combination with other optical modalities to improve ...Missing: papers | Show results with:papers<|control11|><|separator|>
  98. [98]
  99. [99]
  100. [100]
  101. [101]
  102. [102]
  103. [103]
  104. [104]
    Multimodal bioimaging across disciplines and scales - Nature
    Mar 1, 2024 · Multimodal bioimaging is a broad term used to describe experimental workflows that employ two or more different imaging modalities.
  105. [105]
    Machine learning empowered coherent Raman imaging and ...
    Jan 25, 2025 · Here, we present a comprehensive review of the latest advancements in the application of machine learning in the molecular spectroscopic imaging fields.
  106. [106]
    Portable Handheld Probe-Based Confocal Raman System
    Portable Handheld Probe-Based Confocal Raman System: A Versatile Tool for Non-Invasive Objective Assessment of Dermatological Conditions.
  107. [107]
    Toward a portable stimulated Raman scattering system
    May 28, 2025 · This study focuses on the development of a portable SRS system based on a tunable dual-output fiber-based light source which can be used not only for medical ...
  108. [108]
    Advancements and Challenges Shape Raman Solutions for Medical ...
    The evolution of Raman spectrometers and microscopes from bulky, laboratory-bound instruments to compact, portable, and clinically viable tools at the point of ...