Fact-checked by Grok 2 weeks ago

Infrared spectroscopy

Infrared spectroscopy, often abbreviated as IR spectroscopy, is a versatile analytical technique that identifies and characterizes chemical substances by measuring their absorption of infrared radiation, which excites specific vibrational modes within molecules. This absorption occurs at wavelengths corresponding to the energy differences between vibrational ground and excited states, producing a spectrum that serves as a unique "fingerprint" for the sample's molecular composition and functional groups. The core principle of infrared spectroscopy relies on the interaction of infrared light with molecular bonds, where happens only when the radiation's matches the natural vibrational of a bond, such as or motions. These vibrations are quantized and governed by the reduced mass and force constant of the bond, analogous to for springs, with the energy given by E = h\nu, where \nu is the vibrational . For a vibration to be observable (IR-active), it must induce a change in the molecule's ; symmetric vibrations, like the symmetric stretch in CO₂, do not produce such a change and are thus inactive. The technique primarily operates in the mid-infrared region, spanning wavenumbers from approximately 4000 to 400 cm⁻¹ (corresponding to wavelengths of 2.5 to 25 μm), where most organic functional groups exhibit characteristic s. Modern infrared spectrometers predominantly employ Fourier transform infrared (FT-IR) methodology, which uses an interferometer to simultaneously measure all wavelengths, followed by mathematical transformation to generate the spectrum, offering advantages in speed, sensitivity, and resolution over traditional dispersive instruments. Samples can be analyzed in various states—gases, liquids, or solids—using techniques like transmission through thin films, attenuated total reflectance (ATR) for solids and liquids without preparation, or diffuse reflectance for powders. Key spectral regions include the functional group area (above 1500 cm⁻¹) for identifying bonds like O-H (3200–3600 cm⁻¹) or C=O (1650–1750 cm⁻¹), and the fingerprint region (below 1500 cm⁻¹) for unique molecular patterns. Infrared spectroscopy finds widespread applications across disciplines, including and for structural elucidation, pharmaceutical analysis for purity and polymorphism detection, for identification, and for . In forensics, it aids in drug and analysis; in and , it assesses quality and composition non-destructively; and in biomedical research, it examines microbial cells, biofilms, and stress responses. Advanced variants, such as near-field nanoscale IR for high-resolution beyond the diffraction limit, extend its utility to surface and biological .

Basic Principles

Definition and Scope

Infrared spectroscopy is a vibrational spectroscopic technique that measures the absorption, emission, or reflection of infrared radiation by matter, primarily corresponding to transitions between molecular vibrational energy levels. This interaction occurs when infrared photons excite molecular bonds, causing stretches, bends, or other deformations that reveal structural information about the sample. The technique is widely applied in chemistry, materials science, and biology to identify and characterize compounds based on their unique spectral fingerprints. The scope of infrared spectroscopy encompasses the infrared region of the , which lies between visible light and microwaves, with wavelengths ranging from approximately 0.78 μm to 1000 μm. This region is subdivided into near-infrared (, 14000–4000 cm⁻¹ or 0.78–2.5 μm), mid-infrared (, 4000–400 cm⁻¹ or 2.5–25 μm), and far-infrared (, 400–10 cm⁻¹ or 25–1000 μm), where spectra are typically reported in wavenumbers (cm⁻¹) for convenience in relating to . The MIR region is most commonly used for probing fundamental vibrational modes of molecular bonds, while NIR captures overtones and combination bands, and FIR addresses low-energy lattice vibrations in solids. The energies involved (roughly 1–15 kcal/) align closely with those required for molecular bond vibrations, enabling selective excitation without electronic transitions. Key advantages of infrared spectroscopy include its non-destructive nature, allowing samples to be analyzed without alteration or consumption, and its ability to provide rapid identification of functional groups such as carbonyls, hydroxyls, and amines through characteristic absorption bands. This makes it particularly valuable for qualitative and quantitative assessments in diverse fields, from pharmaceutical to forensic analysis.

Historical Development

The discovery of infrared radiation is credited to British astronomer , who in 1800 observed that the temperature of a increased when placed beyond the red end of the produced by passing through a , indicating the presence of invisible radiation with heating effects. This finding laid the groundwork for infrared spectroscopy by establishing the existence of the infrared portion of the . Early efforts to measure infrared spectra were advanced by American physicist William Coblentz, who between 1905 and 1906 systematically recorded infrared absorption spectra of over 100 pure organic and inorganic substances using a detector and rock salt prism spectrometer. His work, published in a seven-part series, provided the first comprehensive catalog of infrared spectra and demonstrated the technique's potential for identifying molecular structures, though limited by low resolution and sensitivity. In the mid-20th century, technological improvements during and after spurred the development of dispersive grating spectrometers, which replaced prisms with diffraction gratings for higher resolution in the and . Commercial instruments became available shortly after the war, with companies like Beckman Instruments and Perkin-Elmer introducing double-beam infrared spectrophotometers in the mid-1940s, enabling routine use in chemical analysis and accelerating adoption in industry and research. The 1960s marked a revolutionary shift with the advent of Fourier transform infrared (FTIR) spectroscopy, leveraging Michelson interferometers to collect interferograms that are mathematically transformed into spectra via fast Fourier transform algorithms, offering superior speed, sensitivity, and resolution over dispersive methods. French physicist Pierre Connes pioneered high-resolution FTIR applications, including planetary atmospheric spectra, while in the United States, Digilab released the first commercial FTIR spectrometer, the FTS-14, in 1970, equipped with minicomputers for data processing. By the 1980s and 1990s, affordable microcomputers and detector arrays made FTIR the dominant technology, vastly expanding its accessibility. In the 2010s, further integration of FTIR with and advanced computing enabled for spatially resolved analysis at the micron scale, enhancing applications in and through automated and multivariate analysis.

Theoretical Foundations

Molecular Vibrations and Modes

Molecules possess vibrational arising from the relative motions of their atoms, which are probed by infrared spectroscopy through changes in moments. For a with N atoms, there are $3N total , accounting for three-dimensional translations and rotations of the atoms. Three of these are translational for the molecule as a whole, and two or three are rotational, leaving the remainder as vibrational . Specifically, non-linear molecules have $3N - 6 vibrational modes, while linear molecules have $3N - 5. Vibrational modes are classified into and types. Stretching modes involve changes in bond lengths and can be symmetric, where bonds lengthen and shorten in phase, or asymmetric, where they move out of phase. modes involve changes in bond angles and include scissoring (atoms approach and separate like scissors), rocking (a group of atoms moves in opposite directions in a plane), wagging (out-of-plane motion perpendicular to the molecular plane), and twisting (rotation about a bond axis). For example, (H₂O), a non-linear with N=3, exhibits three vibrational modes: a symmetric stretch, an asymmetric stretch, and a (scissoring) mode. (CO₂), a linear with N=3, has four vibrational modes: a symmetric stretch, an asymmetric stretch, and two degenerate modes. These vibrational modes are described as normal modes, which represent independent, collective oscillations of the atoms where all parts of the move in phase with the same frequency. In the harmonic oscillator approximation, valid for small-amplitude vibrations, each normal mode behaves like an independent . The vibrational frequency \nu for a mode is approximated by \nu \approx \frac{1}{2\pi} \sqrt{\frac{k}{\mu}}, where k is the force constant of the bond or angle, and \mu is the of the oscillating units. This model assumes quadratic surfaces, simplifying the description of molecular vibrations. Real molecules exhibit due to deviations from the potential, particularly at larger amplitudes, leading to effects such as (multiples of fundamental frequencies) and combination bands in spectra. arises from interactions between modes and finite bond lengths, causing the to include higher-order terms beyond the .

Quantum Mechanical Basis

Infrared absorption arises from the quantization of molecular vibrational levels, which can be modeled using the quantum mechanical for diatomic or polyatomic bonds under the small-displacement approximation. The of this model is parabolic, V(r) = \frac{1}{2} k (r - r_e)^2, where r is the internuclear distance, r_e is the equilibrium distance, and k is the force constant. Solving the yields discrete levels given by E_v = h \nu \left( v + \frac{1}{2} \right), where v = 0, 1, 2, \dots is the vibrational quantum number, h is Planck's constant, and \nu is the classical vibrational frequency, \nu = \frac{1}{2\pi} \sqrt{\frac{k}{\mu}}, with \mu as the reduced mass. In this ideal case, transitions between levels occur only for changes \Delta v = \pm 1, corresponding to fundamental vibrational absorptions. Real molecular bonds deviate from perfect harmonicity due to anharmonic effects, where the potential flattens at larger displacements and includes repulsive walls at short distances. The provides a more accurate description: V(r) = D_e \left( 1 - e^{-a(r - r_e)} \right)^2, with D_e as the from the bottom of the well and a a parameter related to the near . This model yields vibrational energy levels E_v = h \nu \left( v + \frac{1}{2} \right) - \frac{(h \nu)^2}{4 D_e} \left( v + \frac{1}{2} \right)^2 + \ higher-order\ terms, allowing for overtones at approximately $2\nu and combination bands involving multiple modes, though these are weaker than fundamentals. Anharmonicity also leads to dissociation at finite energy, limiting the number of bound states to roughly v_{max} \approx \frac{D_e}{h \nu}. The interaction between molecules and infrared radiation is governed by time-dependent perturbation theory, where the oscillating of the acts as a on the . occurs when the h \nu_{photon} matches the difference \Delta E between vibrational states, promoting the molecule from a lower to a higher level. The rate is derived using first-order time-dependent , yielding the probability per unit time proportional to the square of the matrix element of the . This formalism connects to Einstein's coefficients: the B_{if} (from initial state i to final state f) determines the rate as w_{if} = B_{if} \rho(\nu), where \rho(\nu) is the radiation at \nu, while the B_{fi} = B_{if} and A_{fi} relate via A_{fi} = \frac{8\pi h \nu^3}{c^3} B_{if}. Both A and B depend on the moment |\mu_{if}|^2. Infrared vibrational transitions typically span energies of 0.01–1 eV, corresponding to wavenumbers of 100–4000 cm⁻¹ in the mid-infrared region, where $1\ \text{cm}^{-1} \approx 1.24 \times 10^{-4}\ \text{eV}. This range aligns with the thermal energy at room temperature (~0.025 eV), enabling population of low-lying vibrational states and observable absorptions.

Selection Rules and Intensity

In infrared spectroscopy, a vibrational transition is allowed only if it results in a change in the molecular dipole moment during the vibration. This condition, known as the gross selection rule for electric dipole transitions, requires that the derivative of the dipole moment μ with respect to the normal coordinate Q of the vibration be non-zero, i.e., \partial \mu / \partial Q \neq 0. Without this change, the interaction between the molecular vibration and the oscillating electric field of the infrared radiation is negligible, rendering the mode infrared-inactive. Within the approximation, the specific for vibrational v dictates that only fundamental transitions with \Delta v = \pm 1 are permitted, corresponding to excitations from the (v=0) to the first (v=1) or vice versa. relaxes this rule, allowing transitions with \Delta v > 1 (e.g., \Delta v = 2 for the first ) and bands, though these are typically weaker due to smaller transition moments. considerations, analyzed through theory, further determine activity: a mode is IR-active if its species transforms like one of the components (x, y, or z). For example, in linear CO₂ (D_{\infty h} ), the symmetric stretch (\sigma_g^+) is inactive because it preserves the , while the asymmetric stretch (\sigma_u^+) is active as it induces a temporary . These rules build on the quantum energy levels of molecular vibrations, ensuring only compatible transitions contribute to the spectrum. The intensity of an IR absorption band reflects the strength of the transition and is quantified by the integrated absorbance A, which is proportional to the square of the transition dipole moment matrix element |\langle \psi_f | \hat{\mu} | \psi_i \rangle|^2 multiplied by the population of the initial state, governed by the . Experimentally, this relates to concentration via the Beer-Lambert law: A = \epsilon c l, where \epsilon is the absorptivity (dependent on the transition moment), c is the concentration, and l is the path length. Stronger dipole changes yield higher \epsilon values, leading to more intense peaks, while thermal population effects favor lower-energy fundamentals over hot bands from excited initial states. IR spectral peaks exhibit characteristic shapes due to broadening mechanisms: from finite vibrational lifetimes produces profiles, while inhomogeneous effects like Doppler or environmental variations yield Gaussian profiles; real spectra often show Voigt profiles as convolutions of both. tails arise from lifetime uncertainty (\Delta E \Delta t \approx \hbar), with (FWHM) inversely proportional to lifetime, whereas Gaussian broadening reflects statistical distributions in velocity or . In the fingerprint region (typically below 1500 cm^{-1}), numerous allowed vibrational modes overlap, producing complex, unique patterns that serve as molecular "fingerprints" for identification, as the interplay of active modes encodes structural details without isolated group assignments. This utility stems from the density of IR-active fundamentals, , and interactions in polyatomic molecules, enabling comparison to reference spectra for compound verification.

Instrumentation

Dispersive Spectrometers

Dispersive spectrometers represent the traditional approach to () spectroscopy, employing optical elements to physically separate wavelengths of before detection. These instruments scan across the sequentially, measuring intensity at individual wavelengths to construct the full point by point. Unlike interferometric methods, they rely on via prisms or within a to achieve wavelength selection. The primary components of a dispersive IR spectrometer include a broadband IR radiation source, a monochromator for dispersion, a detector for signal measurement, and a mechanical chopper for signal modulation. Common sources are the Globar, a silicon carbide rod heated electrically to 1300–1500 K, which emits continuous radiation peaking around 2 μm and covering approximately 0.5–50 μm, and the Nernst glower, a cylindrical filament of rare earth oxides (such as , , and oxides) heated to approximately 2000 K, providing emission from 0.5–25 μm with higher intensity in the mid-IR region. The monochromator typically uses either prisms made of alkali halides like NaCl (effective for 2.5–15 μm but limited to resolutions of about 10 cm⁻¹ due to material dispersion) or reflection gratings (blazed for IR wavelengths, capable of resolutions down to ~1 cm⁻¹ across broader ranges). Detectors are thermal devices such as thermocouples, which generate a voltage from the temperature difference between an IR-absorbing hot junction and a reference cold junction, or bolometers, which measure resistance changes in a temperature-sensitive element (e.g., metal film or ) upon IR absorption. A rotating chopper wheel, operating at 5–30 Hz, modulates the IR beam to convert the signal to an alternating current, enabling lock-in amplification to distinguish it from ambient thermal noise. In operation, IR radiation from the source passes through sample and reference compartments in a double-beam configuration to compensate for source fluctuations and solvent absorption. The beams then enter the , where the or disperses the radiation into its spectral components; a scanning mechanism rotates the dispersive element or adjusts entrance/exit slits to select successive narrow bands (typically 1–10 cm⁻¹ wide). The modulated beam at each selected reaches the detector, producing an electrical signal proportional to , which is recorded as the scans from ~4000 cm⁻¹ to 400 cm⁻¹. This sequential acquisition requires sensitive detectors due to the low flux in IR. Dispersive spectrometers offer high spectral resolution in targeted narrow wavelength ranges, making them suitable for detailed studies of specific bands, with grating-based systems achieving ~1 cm⁻¹ . However, their point-by-point scanning leads to slow acquisition times (often several minutes per spectrum) and lower signal-to-noise ratios compared to IR (FTIR) spectrometers, which benefit from multiplex detection. Historically, dispersive instruments dominated IR spectroscopy from their development in the mid-1940s through the , enabling widespread adoption in organic structural analysis, but they were largely supplanted by commercial FTIR systems in the late due to the latter's speed and sensitivity advantages. They remain in use today for specialized high-resolution applications, such as monitoring single IR wavelengths in kinetic studies or near-IR regions.

Fourier Transform Infrared (FTIR) Spectrometers

Fourier transform (FTIR) spectrometers represent a significant advancement in instrumentation, employing to measure spectra in the before computationally converting data to the . At the core of an FTIR spectrometer is the , which splits incoming radiation from a broadband source using a , directing one beam to a fixed mirror and the other to a moving mirror. The recombined beams interfere, producing an interferogram as a function of the difference δ created by the moving mirror's translation. This interferogram I(δ) is mathematically described by the I(\delta) = \int_{-\infty}^{\infty} B(\nu) \cos(2\pi \nu \delta) \, d\nu, where B(ν) is the intensity as a function of ν. To obtain the infrared spectrum, the interferogram undergoes a , yielding the intensity spectrum B(ν) via B(\nu) = 2 \int_{0}^{\infty} I(\delta) \cos(2\pi \nu \delta) \, d\delta. In practice, this is implemented using a () on digitized interferogram data sampled at precise intervals, enabling efficient computation even for high-resolution spectra. Unlike dispersive spectrometers that sequentially wavelengths, FTIR captures the entire simultaneously in each interferogram measurement. FTIR offers key advantages over dispersive methods, including the Fellgett or multiplex advantage, where all wavelengths contribute to the signal throughout the measurement, improving the signal-to-noise ratio (S/N) by a factor approximately proportional to the square root of the number of resolution elements for detector-noise-limited cases. Additionally, the Jacquinot or throughput advantage arises from the interferometer's slitless design, allowing a larger aperture and higher light collection efficiency—often 10 to 100 times greater than in dispersive systems—enhancing sensitivity for low-light samples. Rapid scanning of the moving mirror further enables high-speed data acquisition, with spectra collected in seconds. The primary components of an FTIR spectrometer mirror those of dispersive instruments—such as the infrared source, sample compartment, and detector—but incorporate interferometer-specific elements like the beam splitter and mirrors, typically coated for infrared transmission. A helium-neon (HeNe) laser provides a monochromatic reference beam for precise zero-path-difference tracking, ensuring accurate sampling during mirror motion via interferogram modulation at the laser's frequency. Detectors commonly include deuterated triglycine sulfate (DTGS) pyroelectric sensors for room-temperature operation or mercury cadmium telluride (MCT) semiconductor detectors cooled with liquid nitrogen for faster response and higher sensitivity in the mid-infrared range. To mitigate artifacts in the transformed spectrum, applies a to the interferogram before transformation, tapering the data edges to suppress from finite truncation while broadening the . Common functions include the Happ-Genzel (similar to a cosine bell) for balanced and or the Blackman-Harris for strong sidelobe suppression in . This processing step is essential for producing clean, interpretable spectra without introducing excessive distortion.

Experimental Methods

Sample Preparation Techniques

Sample preparation in () spectroscopy is essential to ensure that the sample interacts effectively with the IR beam while minimizing artifacts and interferences, allowing for accurate or measurements across gas, liquid, and phases. Traditional methods focus on creating thin, uniform sample layers or controlled environments to achieve optimal path lengths and avoid saturation of strong bands. For gas samples, long-path gas cells with path lengths of 10 to 100 cm are commonly employed to enhance sensitivity for low-concentration analytes, typically fitted with NaCl windows that transmit mid- radiation effectively. These cells allow the gas to flow through or be statically held, enabling the recording of absorption spectra as the IR beam passes multiple times via mirrors. Pressure broadening effects must be considered, as increased gas leads to collisional line broadening, which can alter and linewidths proportionally to pressure, often requiring operation at reduced pressures (e.g., 10-100 ) for sharp rotational-vibrational features. Liquid samples are prepared as thin films sandwiched between two IR-transparent plates, such as KBr or NaCl, to achieve a path length of about 0.01-0.05 mm and prevent total absorption in strong bands like C-H stretches. For non-volatile liquids, this method provides a simple setup, though volatile solvents may require sealed cells to avoid evaporation. mulls, involving dispersion in between plates, are occasionally adapted for viscous liquids but are more standard for solids; they introduce C-H bands that must be subtracted. (ATR) techniques bypass extensive preparation for non-volatile liquids by directly placing the sample on a surface, where the evanescent wave probes a shallow depth (typically 0.5-5 μm). Solid samples often require dispersion in an IR-transparent matrix to form uniform mixtures for analysis. The KBr pellet method involves grinding 1-2% by weight of the finely powdered sample ( <2 μm) with KBr, then pressing at 10-15 tons to form a 1 mm thick, 13 mm diameter disk that is nearly transparent to . This technique suits crystalline solids but demands dry conditions to avoid moisture artifacts. For powders and irregular solids, diffuse reflectance infrared (DRIFTS) collects scattered from the sample surface diluted in KBr (up to 5-10%), providing spectra without pressing and suitable for of surface properties. Polymers, which may not disperse well, can be analyzed via , where a small sample (1-10 mg) is heated to 500-800°C in a controlled , volatilizing fragments for gas-phase detection with minimal initial preparation. Common challenges in include interference, where atmospheric moisture or residual H₂O in samples produces broad stretching bands around 3700-3500 cm⁻¹ and at 1640 cm⁻¹, overlapping key regions and necessitating dry purging or desiccated environments. subtraction is critical when using solutions, involving recording a background of the pure and digitally removing its contributions to isolate bands, though incomplete subtraction can introduce baseline distortions. Since the early , diamond ATR crystals have become a dominant modern alternative, offering durable, high-pressure-resistant surfaces for minimal-preparation analysis of all sample types; a small amount of liquid, solid, or powder is simply pressed against the crystal, enabling rapid, non-destructive measurements with penetration depths of 1-3 μm and effective range down to 400 cm⁻¹. This approach has largely supplanted traditional methods in routine labs due to its ease and reproducibility.

Spectral Acquisition and Analysis

Spectral acquisition in infrared (IR) spectroscopy begins with the collection of a background spectrum, also known as an empty cell or reference scan, which captures instrumental and environmental contributions such as solvent or atmospheric interferences. This background is subtracted from the sample spectrum to isolate the analyte's signal, ensuring accurate representation of molecular vibrations. Typical resolution settings range from 4 cm⁻¹ to 8 cm⁻¹, with 4 cm⁻¹ being common for routine analyses to balance detail and acquisition time without excessive broadening of peaks. To enhance the signal-to-noise ratio (S/N), multiple scans—often 16 to 64—are averaged, as the S/N improves proportionally to the square root of the number of scans, reducing random noise while preserving spectral features. Following acquisition, refines the raw spectra for reliable interpretation. Baseline correction eliminates drifting offsets or sloping backgrounds caused by or instrumental drift, using methods such as polynomial fitting or piecewise linear subtraction to flatten the spectrum without distorting peak shapes. Smoothing reduces high-frequency noise via the Savitzky-Golay filter, which performs local least-squares polynomial fitting over a moving window (typically 11–21 points) to preserve peak widths and heights better than simple averaging. Normalization scales the spectrum to a standard intensity, often to unit area or maximum , facilitating comparisons across samples. Spectra are commonly converted from (T = I/I₀) to (A = -log₁₀ T) units, as follows a linear relationship with concentration per Beer's law, aiding quantitative work. Qualitative analysis involves peak picking to identify maxima corresponding to vibrational modes, followed by matching the overall "" region (typically 1500–400 cm⁻¹) against reference libraries. Automated algorithms detect peaks above a , assigning wavenumbers and relative intensities for inference. Databases like the NIST Chemistry WebBook provide evaluated spectra for thousands of compounds, enabling hit-quality indexing via similarity metrics such as correlation coefficients to confirm molecular identities. For , Beer's law (A = ε b c, where ε is the molar absorptivity, b the path length, and c the concentration) underpins concentration determination from peak heights or areas at characteristic wavenumbers. In multicomponent mixtures, least-squares fits the observed spectrum as a of reference spectra, solving for individual concentrations while accounting for overlapping bands. This method assumes additivity of absorbances and is effective for systems obeying the law, with matrix inversion or iterative algorithms minimizing residuals. Common artifacts must be addressed to avoid misinterpretation. Atmospheric CO₂ and H₂O bands (e.g., CO₂ at ~2350 cm⁻¹ and H₂O at ~3400–3600 cm⁻¹ and 1600 cm⁻¹) are removed by subtracting a pre-recorded atmospheric or through purging with dry , ensuring they do not overlap sample features. , which causes baseline elevation and reduced , is corrected by measuring response with a blocked and subtracting this scatter contribution, particularly important in dispersive systems.

Spectral Features

Absorption Regions and Bands

The infrared spectrum is conventionally divided into three primary regions based on or : the near-infrared (, approximately 14,000–4,000 cm⁻¹), mid-infrared (, 4,000–400 cm⁻¹), and far-infrared (, 400–10 cm⁻¹). The MIR region is the most widely utilized for molecular vibrational analysis, as it corresponds closely to the energies of vibrational transitions. Within the MIR, the higher-wavenumber portion from 4,000 to 1,500 cm⁻¹ is known as the group frequency region, where stretching involving light atoms, such as X-H bonds (X = C, N, O), dominate and provide initial clues to presence. The lower MIR portion, from 1,500 to 600 cm⁻¹, constitutes the fingerprint region, characterized by complex skeletal bending, stretching, and deformation modes unique to the overall molecular structure, enabling compound identification through . Below 600 cm⁻¹ in the FIR, absorption bands arise from heavy-atom motions, including metal-ligand in coordination compounds and lattice in ./Spectroscopy/Vibrational_Spectroscopy/Infrared_Spectroscopy/Infrared_Spectroscopy) In the NIR region, absorptions primarily result from and bands of MIR vibrations, such as the first of O-H stretching appearing around 6,800 cm⁻¹ (approximately twice the near 3,400 cm⁻¹). These bands are inherently weaker—often 10 to 100 times less intense than fundamentals—due to the anharmonic of molecular potentials, which reduces probabilities for higher-order processes. Modern extensions into the low-wavenumber FIR and (THz) regime (down to 10 cm⁻¹ or ~0.3 THz) have expanded applications to modes in crystalline materials, revealing low-energy collective vibrations not accessible in standard MIR measurements. IR absorption bands exhibit characteristic shapes and widths influenced by molecular environment and relaxation processes. In condensed phases like liquids or solutions, typical band widths at half-height range from 10 to 20 cm⁻¹, arising from vibrational relaxation (on timescales) and rotational dephasing, which shorten the excited-state lifetime and broaden lines via the energy-time . Asymmetry in band profiles often stems from , where a fundamental couples with a nearby or band of similar energy and , leading to splitting and intensity redistribution. further modulate band characteristics; for instance, hydrogen bonding in protic solvents causes O-H stretching bands to shift to lower wavenumbers (e.g., from ~3,600 cm⁻¹ in dilute non-polar media to ~3,300 cm⁻¹ in ) and broaden significantly due to inhomogeneous broadening from varying hydrogen-bond strengths and rapid ./Spectroscopy/Vibrational_Spectroscopy/Vibrational_Modes/Combination_Bands_Overtones_and_Fermi_Resonances)

Characteristic Group Frequencies

Characteristic group frequencies in infrared spectroscopy correspond to the distinctive ranges associated with the vibrational modes of specific functional groups, enabling the identification of molecular structures through spectral . These absorptions are primarily due to and vibrations of bonds, with modes generally appearing at higher s than deformations. The consistency of these frequencies across similar compounds makes them invaluable for qualitative analysis, often serving as the first step in interpreting an IR spectrum. The diagnostic power lies in correlating observed peaks to known group frequencies, though exact positions can vary slightly based on molecular environment. For example, the C-H stretching vibrations distinguish between aliphatic and unsaturated hydrocarbons: alkanes absorb at 3000-2850 cm⁻¹, while aromatic C-H stretches occur at 3100-3000 cm⁻¹. Similarly, the carbonyl C=O stretch is a hallmark of ketones, aldehydes, and carboxylic derivatives, typically in the 1750-1650 cm⁻¹ range, with subtypes like esters at higher frequencies (~1735 cm⁻¹) and conjugated amides at lower (~1650 cm⁻¹). Hydroxyl O-H stretches provide another clear indicator, with free O-H in dilute solutions showing sharp peaks at 3650-3580 cm⁻¹ and hydrogen-bonded O-H in alcohols or appearing as broad bands from 3550-3200 cm⁻¹. The following table summarizes representative characteristic frequencies for common organic functional groups, focusing on stretching modes for diagnostic purposes:
Functional GroupVibration TypeWavenumber Range (cm⁻¹)Typical IntensityNotes/Example
C-HStretch3000–2850Strong (s)Symmetric and asymmetric stretches in CH₃ and CH₂ groups.
Aromatic C-HStretch3100–3000Medium (m)Often accompanied by =C-H bend at 900–700 cm⁻¹.
O-H (alcohols, ; H-bonded)Stretch3550–3200Strong, broad (s, br)Broadening due to hydrogen bonding; free O-H sharper at ~3600 cm⁻¹.
C=O (ketones)Stretch1720–1700Strong (s)Saturated acyclic; shifts lower with conjugation.
C=O (esters)Stretch1750–1730Strong (s)Higher frequency due to electronegative oxygen.
C=O (amides)Stretch1680–1630Strong (s)Lower due to with N .
C≡C (alkynes)Stretch2260–2190Variable (m-w)Weak if symmetric; stronger in terminal alkynes.
C=C (alkenes)Stretch1680–1620Variable (m-w)Intensifies with conjugation.
These values are drawn from standard correlation charts and may vary by ±10-20 cm⁻¹ depending on the specific compound. Factors such as electronic delocalization and steric constraints can shift these frequencies predictably. Conjugation with adjacent π-systems lowers stretching frequencies by reducing bond strength; for instance, the C=O stretch in α,β-unsaturated ketones or conjugated esters drops by 20-40 cm⁻¹ relative to unconjugated analogs. Conversely, ring strain in cyclic structures increases frequencies by compressing bond angles and strengthening bonds; five-membered lactones exhibit C=O stretches ~30 cm⁻¹ higher than acyclic esters, and cyclopropanone derivatives can reach 1800 cm⁻¹. Below 1500 cm⁻¹ lies the fingerprint region, where overlapping absorptions from skeletal vibrations and low-energy modes create a unique pattern for each molecule, akin to a spectral "fingerprint." This area is less useful for group identification but essential for confirming compound identity by direct comparison to reference spectra, as no two isomers produce identical patterns here. Environmental conditions further modulate these frequencies, emphasizing the need for standardized measurement. In carboxylic acids, intermolecular hydrogen bonding leads to dimer formation, broadening and shifting the O-H stretch to 3000-2500 cm⁻¹ with a characteristic "twin peak" profile, while monomeric forms in dilute non-polar solvents show sharper bands near 3550 cm⁻¹. Temperature increases can weaken hydrogen bonds, sharpening and shifting O-H bands to higher wavenumbers, and pH variations in ionic species alter vibrations; for example, deprotonated carboxylates (COO⁻) display asymmetric and symmetric stretches at ~1600 and ~1400 cm⁻¹ instead of the acidic O-H. For applications in modern , characteristic frequencies extend to inorganic and organometallic groups. Silicon-based functionalities, prevalent in polymers and semiconductors, show Si-H stretches at 2250-2100 cm⁻¹ in silanes and polysilanes, useful for monitoring in silicon surfaces. Si-O stretches in siloxanes and silicates appear at 1100-1000 cm⁻¹, often as strong, broad bands indicative of network formation. In organometallic compounds, (M-CO) stretches serve as probes for , with terminal CO ligands absorbing at 2100-1850 cm⁻¹ (higher for less electron-donating metals) and bridging CO at lower frequencies (~1800-1700 cm⁻¹), enabling distinction between mononuclear and polynuclear complexes.

Badger's Rule and Bond Strength

Badger's rule provides an empirical correlation between the bond length r_\mathrm{eq} and the stretching force constant k of diatomic molecules, expressed approximately as k = a (1/r_\mathrm{eq})^3, where a is an atom-specific constant derived from periodic table row dependencies. Formulated in 1935, this rule captures how shorter bonds correspond to stronger vibrational potentials, enabling the prediction of bond properties from spectroscopic data. In infrared spectroscopy, the observed vibrational frequency \bar{\nu} (in cm^{-1}) for a diatomic stretch relates directly to the force constant via the harmonic oscillator approximation: \bar{\nu} = \frac{1}{2\pi c} \sqrt{\frac{k}{\mu}}, where c is the and \mu is the of the atoms. Consequently, stronger bonds with larger k (e.g., triple bonds versus single bonds) exhibit higher stretching frequencies; for instance, C≡N stretches near 2200 cm^{-1}, while C-N appears around 1000 cm^{-1}, reflecting the inverse relationship between , length, and \bar{\nu}. This connection allows IR spectra to qualitatively assess bond strengths without direct structural determination./Spectroscopy/Vibrational_Spectroscopy/Infrared_Spectroscopy/Infrared_Spectroscopy) The finds practical application in predicting vibrational frequencies from molecular structures or estimating bond lengths from measured IR frequencies, particularly useful for diatomic-like stretches in larger systems. However, it is limited to harmonic approximations and primarily validated for diatomics, showing deviations in anharmonic or highly coupled vibrations. Extensions like the Urey-Bradley address polyatomic molecules by incorporating central (non-bonded) repulsive terms F for 1,3 interactions alongside bond stretches and angle bends, improving fits to observed IR and Raman spectra. Post-2010 computational studies using have examined Badger's rule across diverse bonds, identifying exceptions while confirming its utility in many cases. For example, a 2015 study on N-F, O-F, N-Cl, P-F, and As-F bonds found deviations from the inverse relationship for certain N-F and O-F bonds in fluoro amines, radicals, and oxides, attributing anomalies to withdrawal, orbital contraction, and delocalization effects. These advancements integrate Badger's rule into quantum chemical workflows for IR spectral assignment, linking empirical observations to theoretical bond strengths.

Advanced Phenomena

Isotope Effects on Spectra

Isotopic substitution alters the vibrational frequencies observed in infrared (IR) spectra by changing the reduced mass of the atoms involved in a molecular vibration, while the force constant remains largely unchanged. For a diatomic oscillator model, the vibrational frequency \nu is given by \nu = \frac{1}{2\pi c} \sqrt{\frac{k}{\mu}}, where k is the force constant, c is the speed of light, and \mu is the reduced mass. Substituting a heavier isotope increases \mu, thereby decreasing \nu. The relative frequency shift can be approximated as \frac{\Delta \nu}{\nu} \approx 1 - \sqrt{\frac{\mu'}{\mu}}, where \mu' is the reduced mass with the heavier isotope. A prominent example is the substitution of hydrogen with deuterium in water molecules. The broad O-H stretching band in H₂O appears around 3400 cm⁻¹, reflecting hydrogen-bonded networks, whereas the corresponding O-D stretch in D₂O shifts to approximately 2500 cm⁻¹ due to the doubled mass of deuterium, which scales the frequency by roughly \sqrt{1/2} \approx 0.707. This shift is well-documented in studies of liquid water isotopomers, where HDO exhibits intermediate frequencies, aiding in the assignment of overlapping modes. Similarly, in carbonyl groups (C=O), replacing ¹²C with ¹³C lowers the stretching frequency by about 40 cm⁻¹, as seen in isotopically labeled peptides where the band shifts from ~1650 cm⁻¹ to ~1610 cm⁻¹. For oxygen, substituting ¹⁶O with ¹⁸O in the same C=O moiety causes a shift of approximately -28 cm⁻¹, providing a distinct marker for oxygen involvement. These isotope-induced shifts are analytically powerful for confirming positions in molecules. By selectively labeling suspected atoms and observing displacements, researchers can assign specific to particular bonds, resolving ambiguities in spectral interpretation. In mechanistic studies of chemical reactions, reveals bond-breaking or formation pathways; for instance, a shift in reaction intermediates indicates transfer involvement. In biomolecules like proteins, multiple isotopic substitutions create combinatorial patterns, enabling site-specific resolution of secondary structures via isotope-edited spectroscopy. This approach has been instrumental in determining orientations and folding dynamics, where shifts isolate individual otherwise overlapped in the 1600–1700 cm⁻¹ region.

Two-Dimensional Infrared Spectroscopy

Two-dimensional infrared (2D-IR) spectroscopy is a nonlinear ultrafast technique that maps vibrational transitions in a two-dimensional , with along one axis and detection along the other. Diagonal peaks in the 2D spectrum correspond to the linear of individual vibrational modes, while off-diagonal cross-peaks arise from anharmonic s or between modes, revealing interactions such as through-bond or through-space due to delocalized excitons. This approach exploits the cubic nonlinearity of the molecular response to sequences of intense pulses, enabling the isolation of third-order signals that encode structural and dynamic information. The presence of cross-peaks is a direct signature of vibrational , which shifts the excited-state relative to the ground-state , typically by a few cm⁻¹ for or carbonyl stretches. The experimental setup relies on mid-IR pulses (∼150 duration, 5–10 μm ) derived from optical amplifiers seeded by Ti:sapphire regenerative amplifiers operating at kHz repetition rates. In the prevalent photon echo configuration, three collinear or noncollinear pulses interact with the sample: the first induces a between ground and excited states, the second transfers it to a between excited and doubly excited states (or creates a ), and the third serves as a for detection of the phase-coherent emitted field. The waiting time between the second and third pulses, tunable from 0 to several nanoseconds, captures relaxation dynamics, while population times between the first and second pulses probe evolution. Data are Fourier-transformed along and detection times to yield the , with phase-sensitive detection preserving sign information for positive (ground-state bleach and ) and negative (excited-state absorption) features. Key variants include the fully rephasing photon echo, which corrects for inhomogeneous broadening by reversing during the waiting time, and the non-rephasing pump-probe method, a simpler using a pump and probe that emphasizes but includes unwanted coherent artifacts. Double-quantum 2D-IR extends this by exciting transitions, enhancing sensitivity to bilinear couplings in congested spectra. These methods are implemented with shapers for control and diffraction gratings for spectral dispersion. Applications of 2D-IR center on resolving vibrational couplings in complex systems, such as protein secondary structures via amide I (∼1600–1700 cm⁻¹) cross-peaks that distinguish α-helices (strong diagonal, weak cross-peaks) from β-sheets (off-diagonal features indicating extended H-bond networks). In studies, time-dependent cross-peak evolution tracks conformational interconversions on 100 ps to μs scales, as demonstrated for villin headpiece folding intermediates. dynamics are probed through frequency-frequency correlation functions derived from diagonal peak linewidths, quantifying electrostatic fluctuations around probes like CO or N₃⁻ groups in (∼1–3 ps decoherence times). In , 2D-IR maps excitonic delocalization in conjugated polymers. Over one-dimensional IR, 2D-IR disperses overlapping bands across two dimensions for improved , quantifies anharmonicities (Δω ≈ 10–20 cm⁻¹) and strengths (off-diagonal intensities), and measures ultrafast timescales (femtoseconds for coherences, picoseconds for populations, nanoseconds for orientations) inaccessible to linear methods, enabling direct observation of energy flow and structural fluctuations. Signal-to-noise requires ∼10⁹–10¹² molecules, limiting it to concentrated samples but offering atomic-scale insights into transient states. Emerging developments in the 2020s include multidimensional extensions like 2D IR-Raman that correlate vibrational and rotational motions for tetrahedral order in liquids, while portable implementations with mid-IR fiber lasers broaden accessibility for studies. These advances enhance sensitivity and multidimensionality for hybrid materials and biological interfaces.

Applications

Chemical Analysis and

Infrared spectroscopy plays a pivotal role in chemical analysis by enabling the of unknown compounds through the of their spectra with established reference . One of the most comprehensive collections is Wiley's IR Spectral Library, which includes the historic Sadtler database originally compiled in the mid-20th century and now encompasses over 339,000 infrared spectra, which facilitates rapid matching for organic and inorganic substances across various states of matter. This library search approach is particularly effective for confirming molecular identities in routine settings, where spectral similarities in the fingerprint region (typically 400–1500 cm⁻¹) allow for high-confidence identifications. Complementing library matching, analysis interprets characteristic bands—such as the C=O stretch around 1700 cm⁻¹ for carbonyls—to elucidate structural features, providing insights into molecular skeletons without requiring crystalline samples. For purity assessment, infrared spectroscopy excels at detecting impurities through the observation of extraneous peaks in the spectrum of a purportedly pure . In pharmaceutical synthesis, for instance, trace contaminants can manifest as weak absorptions distinct from the main product's bands, allowing quantification via peak intensity ratios when calibrated against standards. Reaction monitoring further leverages this capability; during esterification processes, the growth of the C=O band at approximately 1735 cm⁻¹ and diminution of the O-H stretch near 3000 cm⁻¹ provide real-time indicators of conversion efficiency and byproduct formation. This analysis minimizes the need for destructive sampling, enhancing throughput in workflows. In the analysis of complex mixtures, hyphenated techniques integrate infrared spectroscopy with chromatographic separations to deconvolute overlapping components. Gas chromatography-infrared (GC-) couples gas-phase separation with vapor-phase IR detection, enabling the identification of volatile organics by eluting fractions into an IR flow cell, where each peak's spectrum is recorded sequentially. Similarly, liquid chromatography-infrared (LC-) hyphenation, often using Fourier-transform IR, handles non-volatile mixtures by interfacing HPLC effluents with a suitable deposition interface or flow cell, thus providing structural confirmation for separated analytes in solution-phase samples. These methods are invaluable for , where they resolve and identify multiple reaction products in a single run. In pharmaceutical applications, infrared spectroscopy is routinely employed for polymorph , as different crystalline forms of the same exhibit subtle shifts in band positions due to variations in hydrogen bonding and packing. For example, the analysis of acetaminophen polymorphs reveals distinct O-H stretching regions around 3200–3400 cm⁻¹, aiding in to ensure formulation stability and . In forensics, portable infrared instruments detect and explosives by matching field-acquired spectra to libraries; nitro-based explosives like show characteristic NO₂ asymmetric stretches near 1530 cm⁻¹, while displays ester carbonyl bands around 1710–1730 cm⁻¹, enabling on-site without sample destruction. Despite its strengths, infrared spectroscopy has limitations in chemical analysis, particularly its inability to distinguish between homopolymers, which produce nearly identical spectra due to shared repeating units and lacking differentiation in vibrational modes. Consequently, it is often complemented by (NMR) spectroscopy for constitutional isomer resolution and (MS) for molecular weight confirmation, forming a multimodal approach that addresses IR's sensitivity to functional groups over full structural detail.

Material and Surface Characterization

Infrared spectroscopy plays a crucial role in characterizing the structural and chemical properties of materials, particularly polymers, surfaces, and , by probing vibrational modes that reveal crystallinity, , , and adsorbate interactions. For s, IR spectra provide insights into crystallinity through the analysis of specific absorption bands, such as the CH₂ rocking modes in (PE), where the ratio of intensities at 730 cm⁻¹ (crystalline) to 720 cm⁻¹ (amorphous) serves as a quantitative measure of the degree of crystallinity. This band ratio approach, often applied in diffuse reflectance infrared (DRIFTS), enables non-destructive evaluation of powder or rough polymer samples by collecting scattered light from the surface, minimizing preparation artifacts and highlighting subtle structural variations. Additionally, IR tracking of monitors changes in functional groups, such as the formation of carbonyls around 1710 cm⁻¹ during oxidative breakdown or the decrease in C-H stretching intensities, allowing real-time assessment of thermal or environmental stability without altering the sample. Surface characterization benefits from specialized IR techniques like reflection-absorption infrared spectroscopy (RAIRS), which enhances sensitivity for thin films and monolayers on reflective substrates by exploiting the metal surface , where p-polarized at grazing incidence amplifies dipole changes perpendicular to the surface. For metallic surfaces, polarization modulation infrared reflection-absorption spectroscopy (PMIRRAS) further improves signal-to-noise ratios by alternating polarization states, enabling precise detection of adsorbate orientations and bonding in monolayers, such as self-assembled thiols on . Hybrid approaches combining surface-enhanced infrared absorption (SEIRA) with plasmonic nanostructures boost signals by factors up to 10⁴ through electromagnetic field enhancement near metal nanoparticles, facilitating the study of weak surface interactions in catalytic systems where adsorbate binding modes, like CO stretching at 2000-2100 cm⁻¹, indicate active site geometries. In , IR spectroscopy elucidates size-dependent vibrational shifts, as quantum confinement alters modes; for instance, in III-V quantum dots, decreasing from 5 nm to 2 nm blueshifts LO frequencies by up to 10 cm⁻¹ due to increased surface-to-volume ratios and strain effects. This is particularly evident in colloidal quantum dots, where Fourier-transform IR reveals confinement-induced changes in vibrations and intra-band transitions, aiding size tuning for optoelectronic applications. For thin films of , such as oriented carbon nanotubes or nanocomposites, IR dichroism—via polarized measurements—quantifies molecular by comparing parallel and perpendicular absorption intensities, with ratios exceeding 2 indicating preferential along the film plane. In , doping effects manifest as shifts in free-carrier absorption bands in the mid-IR (e.g., 1000-2000 cm⁻¹ for n-type ), where increased dopant concentrations like broaden and redshift these plasmons, correlating with carrier densities up to 10¹⁹ cm⁻³. These applications underscore IR's versatility in probing material interfaces and nanostructures, bridging bulk properties with surface-specific behaviors.

Biomedical and Environmental Uses

Infrared spectroscopy plays a pivotal role in biomedical applications, particularly in tissue imaging for cancer diagnostics, where infrared (FTIR) spectroscopy leverages characteristic absorption bands such as I (around 1650 cm⁻¹, associated with C=O ) and II (around 1550 cm⁻¹, linked to N-H bending and C-N ) to differentiate malignant from healthy tissues based on protein conformational changes. This approach enables label-free, non-destructive analysis of histopathological samples, with studies demonstrating high sensitivity in identifying tumor margins in colorectal and other cancers through spectral variations in and content. FTIR further enhances this by providing spatial resolution for , allowing automated guidance of without prior staining, as shown in and tissue analyses. Additionally, near- (NIR) spectroscopy facilitates non-invasive by detecting overtone vibrations of O-H bonds in glucose molecules around 1600-1700 nm, offering real-time assessment with portable devices and minimizing the need for invasive finger pricks. In protein studies, mid- spectroscopy determines secondary structures like α-helices and β-sheets via I band deconvolution, providing insights into folding and aggregation in biological systems. In environmental monitoring, infrared spectroscopy excels in gas analysis for pollutants such as CO₂ and , where open-path FTIR systems measure atmospheric concentrations over long distances by detecting lines in the mid-infrared region (e.g., CO₂ at 2350 cm⁻¹ and NOx around 1300-1900 cm⁻¹), enabling real-time quantification of emissions from industrial sources and vehicles. applications extend this to detection, with airborne imaging spectrometers like AVIRIS-NG mapping and CO₂ plumes from point sources such as oil fields, achieving detection limits below 10 ppm-m through hyperspectral analysis. For water quality assessment, FTIR spectroscopy identifies organic pollutants like and hydrocarbons in by their spectra in the 1000-1700 cm⁻¹ range, supporting rapid screening in electronic industry effluents and riverine systems without extensive . The potential of infrared spectroscopy in biomedical contexts arises from its non-invasive nature, allowing real-time tissue interrogation with minimal sample disruption, while portable FTIR units facilitate field-deployable , reducing logistical challenges in remote pollution assessments. However, challenges persist, including strong water absorption in biological samples that overlaps with key bands (e.g., O-H stretching at 3300 cm⁻¹), necessitating techniques like (ATR) for mitigation, and the requirement for robust models to ensure quantitative accuracy amid matrix variability.

Emerging Techniques

Computational Infrared Spectroscopy

Computational infrared spectroscopy employs quantum chemical methods to predict and simulate infrared (IR) spectra, enabling the of molecular without experimental measurements. These simulations are particularly valuable for understanding complex systems where experimental data may be challenging to obtain, such as gas-phase conformers or transient species. By calculating vibrational frequencies and intensities, researchers can assign spectral features to specific molecular motions, aiding in structural elucidation. Density functional theory (DFT), especially with the B3LYP functional, is widely used for computing vibrational frequencies due to its balance of accuracy and computational efficiency. Frequencies obtained from B3LYP calculations typically overestimate experimental values, necessitating empirical scaling factors of approximately 0.96 to align predicted spectra with observed ones. These scaling factors account for systematic errors in the harmonic approximation and have been derived from extensive benchmark studies across diverse molecular datasets. In the harmonic approximation, vibrational frequencies are derived from the diagonalization of the , which represents the second derivatives of the with respect to coordinates at the equilibrium . Infrared intensities are determined from the derivatives of the molecular with respect to these coordinates, quantifying the change in during . This approach assumes small oscillations around the minimum, providing a foundation for analysis. Popular software packages like Gaussian and facilitate these calculations, offering tools for optimization, computation, and of vibrational modes through animated displacements. Applications of computational IR spectroscopy include conformational analysis, where predicted spectra distinguish between low-energy isomers by matching calculated band positions and intensities to experimental data. For instance, in peptide studies, DFT simulations help identify dominant conformers in isolated gas-phase environments. Additionally, these methods support virtual screening in drug discovery by generating IR fingerprints for large compound libraries, allowing rapid comparison with reference spectra for identification or similarity assessment. Recent advances address limitations of the approximation through anharmonic methods like vibrational configuration interaction (VCI), which incorporates cubic and quartic potential terms to capture mode couplings and for higher accuracy. VCI calculations, often built on DFT potentials, have improved predictions for polyatomic molecules, reducing errors in frequencies by up to 20 cm⁻¹ compared to results. In the 2020s, (ML) techniques have accelerated IR spectrum predictions, with models trained on quantum chemical data enabling near-DFT accuracy at semi-empirical speeds for high-throughput applications. For example, foundation models like MACE4IR use equivariant neural networks to simulate spectra for complex systems efficiently.

Portable and Imaging Methods

Portable infrared spectroscopy has advanced significantly with the development of handheld Fourier-transform (FTIR) spectrometers designed for field applications, such as hazardous materials (HazMat) detection and on-site chemical identification. These devices are typically battery-powered, lightweight (around 2 kg), and offer spectral resolutions of approximately 4-8 cm⁻¹, enabling rapid analysis of solids, liquids, and powders without extensive sample preparation. For instance, instruments like the Agilent 4300 Handheld FTIR and the 908 Devices ProtectIR allow and to identify chemical threats in seconds by using (ATR) sampling interfaces. Their portability stems from miniaturized Michelson interferometers and robust enclosures that withstand environmental stresses, making them ideal for non-laboratory settings. Infrared imaging techniques leverage focal plane array (FPA) detectors to generate hyperspectral maps, providing both spatial and chemical information simultaneously across large sample areas. FPAs, consisting of thousands of mercury cadmium telluride (MCT) or other infrared-sensitive pixels, enable snapshot imaging where each pixel records a full spectrum, achieving resolutions down to diffraction-limited scales of 5-10 µm in the mid-IR range. Systems like the Bruker Hyperion microscope with FPA detectors facilitate the visualization of molecular distributions in heterogeneous materials, such as polymer blends or biological tissues, by producing false-color images based on spectral signatures. Synchrotron-based IR imaging further enhances resolution, utilizing the high brightness of synchrotron sources (100-1000 times greater than globar lamps) to achieve spatial resolutions as fine as 2-3 µm, which is crucial for nanoscale chemical mapping in fields like materials science. This approach combines multiple synchrotron beams with wide-field FPA detection for rapid, high-fidelity hyperspectral data acquisition. IR microscopy employs specialized , such as Cassegrain objectives, to focus infrared beams onto small sample areas for detailed and . These reflective objectives, composed of concentric mirrors, correct for chromatic aberrations in the IR range and enable non-contact or contact-mode imaging with magnifications up to 36×, allowing the detection of chemical heterogeneity at micrometer scales. In pharmaceutical applications, IR microscopy reveals distribution variations of active ingredients and excipients within tablets, aiding by identifying uneven coatings or polymorphic forms without destructive sampling. Automated stage scanning with stationary Cassegrain builds comprehensive chemical images, highlighting domains of differing composition that influence drug dissolution. Variant techniques extend IR capabilities to challenging samples. Photoacoustic IR (PA-IR) spectroscopy detects acoustic waves generated by modulated IR absorption, making it suitable for opaque or highly scattering materials where transmission methods fail. This method requires no sample preparation and provides depth profiling up to several millimeters, as the photoacoustic signal arises from thermal expansion in the sample volume. Terahertz (THz) imaging serves as a low-energy extension of IR spectroscopy (0.1-10 THz, or 3-300 cm⁻¹), penetrating non-conductive opaque materials like plastics and ceramics for non-destructive inspection, with applications in security screening and pharmaceutical packaging analysis. THz systems often use time-domain spectroscopy for high-resolution imaging, revealing hidden defects or contraband. Recent innovations include drone-mounted IR spectrometers for large-scale environmental surveys, enabling remote detection of pollutants like leaks over expansive areas. These systems integrate compact FTIR or laser-based IR sensors with unmanned aerial vehicles (UAVs), achieving sensitivities down to parts-per-million levels while covering kilometers in a single flight. For example, broadband FTIR setups on drones perform multi-species gas sensing for atmospheric monitoring, supporting rapid response to emissions from oil fields or landfills. This aerial approach minimizes human exposure to hazardous sites and provides geospatial data for environmental impact assessments.

References

  1. [1]
    Infrared Spectroscopy - MSU chemistry
    Infrared spectrometers, similar in principle to the UV-Visible spectrometer described elsewhere, permit chemists to obtain absorption spectra of compounds.Missing: applications | Show results with:applications
  2. [2]
    [PDF] INFRARED SPECTROSCOPY (IR)
    The IR spectrum is basically a plot of transmitted (or absorbed) frequencies vs. intensity of the transmission (or absorption). Frequencies appear in the x-axis ...
  3. [3]
    IR Spectroscopy - Electronic Materials Research Group - MIT
    Perhaps the most widely used applied spectroscopic method, infrared (IR) absorption spectroscopy probes the characteristic vibration spectra of molecules, and ...
  4. [4]
    [PDF] FT-IR vs. Raman Spectroscopy - NUANCE - Northwestern
    •M-O in the wavenumber range of 400-600 cm⁻ı. Regions in IR Spectrum. 4000 to 400 cm-1. Functional group. Finger print. Above 1500 cm-1. Below 1500 cm-1.
  5. [5]
    IR Spectrum and Characteristic Absorption Bands – Organic Chemistry
    The wavenumber is defined as the reciprocal of wavelength, and the wavenumbers of infrared radiation are normally in the range of 4000 cm-1 to 600 cm-1.
  6. [6]
    Infrared Spectroscopy - Penn State Materials Research Institute
    Infrared spectrometry identifies organic and inorganic compounds, and functional groups. FT-IR analyzes solids, including fibers, films, and powders.
  7. [7]
    Infrared Imaging and Spectroscopy Beyond the Diffraction Limit | NIST
    Jul 22, 2015 · This review presents some basics of infrared spectroscopy and microscopy, followed by detailed descriptions of s-SNOM and PTIR working principles.
  8. [8]
    Infrared Spectrometry - MSU chemistry
    Transitions between vibrational energy states may be induced by absorption of infrared radiation, having photons of the appropriate energy.
  9. [9]
    Spectroscopy: A Measurement Powerhouse | NIST
    Mar 28, 2025 · In absorption spectroscopy, scientists measure light that has passed through clouds of atoms or molecules. Each type of atom or molecule absorbs ...Missing: principles | Show results with:principles
  10. [10]
    Electromagnetic (EM) Spectrum - UCAR Center for Science Education
    The Invisible Portions of the Electromagnetic Spectrum​​ Infrared radiation has wavelengths from 780 nm to 1,000,000 nm (or 1 mm), longer than those of visible ...<|control11|><|separator|>
  11. [11]
    A Review of Mid-Infrared and Near-Infrared Imaging - NIH
    The IR region in electromagnetic spectrum is divided into three regions. These areas are defined as near infrared (near-IR, 13,500–4000 cm−1, 780–2500 nm), mid ...
  12. [12]
    Leveraging infrared spectroscopy for automated structure elucidation
    Nov 16, 2024 · In contrast, IR spectroscopy is quick, cost-effective, non-destructive, and easy to use.
  13. [13]
    Herschel's Experiment - Cool Cosmos - Caltech
    What Herschel had discovered was a form of light (or radiation) beyond red light, now known as infrared radiation.
  14. [14]
    [PDF] Infrared Spectroscopy in Conservation Science - Authentication in Art
    Apr 5, 2025 · IR region into the near (NIR), middle (IR or mid-IR), and far (FIR) sub regions. The majority of analytical applications are found in the middle.
  15. [15]
    Infrared Spectroscopy
    Oct 29, 2020 · The early to mid 1940s saw the first commercial instruments come on the market from both Beckman and Perkin Elmer. In 1957, Perkin Elmer ...Missing: post | Show results with:post
  16. [16]
  17. [17]
    Infrared Spectroscopic Imaging Advances as an Analytical ...
    Modern IR spectroscopic imaging can be traced back nearly 25 years, with the coupling of an IR microscope to an array detector and an interferometer.Missing: FTIR | Show results with:FTIR
  18. [18]
    Number of Vibrational Modes in a Molecule - Chemistry LibreTexts
    Jan 29, 2023 · The normal modes of vibration are: asymmetric, symmetric, wagging, twisting, scissoring, and rocking for polyatomic molecules.
  19. [19]
    Infrared Spectroscopy - Chemistry LibreTexts
    Jan 29, 2023 · Four bending vibrations exist namely, wagging, twisting, rocking and scissoring. ... CO2 has 2 stretching modes, symmetric and asymmetric.
  20. [20]
    Infrared And Raman Spectra Of Polyatomic Molecules
    Jan 22, 2017 · Infrared And Raman Spectra Of Polyatomic Molecules. by: D Van ... PDF WITH TEXT download · download 1 file · SINGLE PAGE PROCESSED JP2 ZIP ...
  21. [21]
    [PDF] Infrared Spectroscopy: Theory
    For noninteracting molecules consisting of N atoms there are 3N 6 vibrational degrees of freedom (3N 5 for a linear molecule), which is equivalent to ...
  22. [22]
    5.5: The Harmonic Oscillator and Infrared Spectra
    Jun 30, 2023 · In the harmonic oscillator model infrared spectra are very simple; only the fundamental transitions, Δ = ± 1 , are allowed.
  23. [23]
    14: Harmonic Oscillators and IR Spectroscopy - Chemistry LibreTexts
    Dec 5, 2019 · The Morse potential, named after physicist Philip M. Morse, is a convenient interatomic interaction model for the potential energy of a diatomic ...
  24. [24]
    [PDF] Experiment S2 INFRARED SPECTROSCOPY Chemistry 114 ...
    That is to say, the Morse potential has the correct shape to fit the observed lower vibrational levels in real molecules, at least to a good approximation.
  25. [25]
    [PDF] Vibration-Rotation Spectra of Diatomic Molecules - CalTech GPS
    Jan 19, 2018 · The great utility of the Morse potential lies in the physical significance associated with each of the terms, and the way it explicitly links ...
  26. [26]
    [PDF] Time-Dependent Perturbation Theory and Molecular Spectroscopy
    Both the Einstein A and B coefficients depend on the square of the matrix element of the dipole-moment operator, given by Eq. (10). We can distinguish two ...
  27. [27]
    [PDF] Quantum Physics III Chapter 4: Time Dependent Perturbation Theory
    Apr 4, 2022 · Energy is transferred from the perturbation to the system, and we have a process of “absorption”. Both cases are of interest. Let us do the ...
  28. [28]
    [PDF] Spectroscopy 1: rotational and vibrational spectra
    Selection rules. The gross selection rule – the electric dipole moment of the molecule must change when the atoms are displaced relative to one another. Such ...
  29. [29]
    [PDF] LASER SPECTROSCOPY OF RADICALS CONTAINING GROUP IIIA ...
    This requirement leads to the selection rules Δv = 0, ±1, ±2, … for totally symmetric modes and Δv = 0, ±2, ±4, … for non-totally symmetric modes. The ...
  30. [30]
    99.05.07: Infrared Spectroscopy: A Key to Organic Structure
    Infrared Spectroscopy: A Key to Organic Structure. Michele Sherban-Kline ... Vibrational transitions that do not result in a change of dipole moment of ...Missing: definition | Show results with:definition
  31. [31]
    [PDF] 2-dimensional infrared spectroscopy and its application to the
    Feb 6, 2014 · A classical treatment of a harmonic oscillator yields a Lorentzian as an approximation to the solution of the system differential equation, but ...
  32. [32]
  33. [33]
    introduction to dispersive and interferometric infrared spectroscopy
    A dispersive spectrometer uses a grating or a prism to disperse radiation from the source into spectral elements, and a detector measures the energy of each ...
  34. [34]
    Infra red Absorption Spectroscopy - Instrumentation
    The heated material will then emit infra red radiation. The Nernst glower is a cylinder (1-2 mm diameter, approximately 20 mm long) of rare earth oxides.Missing: historical development until 1970s cm-
  35. [35]
    [PDF] Dispersive IR Spectroscopy
    • Mid-IR light: 400-4000 cm-1 (25 – 2.5 μm wavelength). • Sources: – Blackbody emitters. • Globar metal oxides. • Nernst Glower: Silicon Carbide. • Detectors ...Missing: historical until 1970s gratings
  36. [36]
    None
    ### Summary of Dispersive IR Spectrometers from Lecture 8 (Chem 4631)
  37. [37]
    IR Thermometers & Pyrometers - Omega Engineering
    Bolometers are essentially resistance thermometers arranged for response to radiation. A sensing element with a thermistor, metal film, or metal wire transducer ...
  38. [38]
    Principles of infrared spectroscopy (2) History of IR spectrometers
    Sep 16, 2022 · Infrared spectrometers were developed in the US in the mid 1940s. Initially, their applications were confined to R&D work on organic compounds.
  39. [39]
    [PDF] Fourier Transform Infrared (FTIR) Spectroscopy
    The Fourier Transform and its Inverse: Allow transformation from the time domain (t) to the frequency domain (ω) and back. Inverse Fourier Transform.<|control11|><|separator|>
  40. [40]
    Fellgett's Advantage in uv-VIS Multiplex Spectroscopy
    The effect of multiplexing in Fourier transform spectroscopy (FTS) (and other coding techniques such as Hadamard transform spectroscopy) on the signal/noise ...
  41. [41]
    Principles of infrared spectroscopy (4) Advantages of FTIR ...
    Sep 16, 2022 · FTIR uses an interferometer, has improved optical throughput, obtains data at multiple wavelengths, improved resolution, and extended ...Missing: components chopper operation disadvantages
  42. [42]
    Upper Boundaries to the Extent of the Jacquinot or Throughput ...
    The typical Fourier transform infrared (FT-ir) spectrometer has a huge light gathering power advantage over grating spectrometers of similar resolution.Missing: FTIR | Show results with:FTIR
  43. [43]
    Infrared Spectroscopy
    IR spectroscopy is a unique instrument in that it can measure any phase of matter. It is also very adaptable to various measurement geometries.
  44. [44]
    FT-IR Spectrometer Detectors - Newport
    Mercury Cadmium Telluride (MCT) detectors exhibit fast response times, a wide dynamic range, and broad spectral response. A two-stage, low noise amplifier is ...
  45. [45]
    Fourier Transform and Apodization - Shimadzu
    Fourier transform is the process of calculating the wave intensity at each period from the sum at all wave periods.
  46. [46]
    Norton-Beer apodization and its Fourier transform
    Oct 17, 2023 · In Fourier transform spectroscopy, apodization is used to alter the instrument line shape, reducing the prominence of its side lobes.
  47. [47]
    4.2: IR Spectroscopy - Chemistry LibreTexts
    Sep 5, 2022 · To prepare a sample for IR analysis using a salt plate, first decide what segment of the frequency band should be studied, refer to Table ...
  48. [48]
    Gemini FTIR Gas Analysis Cells | International Crystal Laboratories
    Long path cells address the most demanding gas sampling requirements of industrial and field deployed FTIR and IR instrumentation. For applications requiring ...Missing: NaCl broadening
  49. [49]
    Fourier transform infrared spectroscopic technique for analysis ... - NIH
    In this technique, a gas sample is put into the cell where it passes through infrared light, and its spectrum can be recorded. The gas cell is typically ...
  50. [50]
    NIR gas phase spectroscopy – Pressure broadening effects
    Oct 4, 2018 · Another important aspect of gas phase spectroscopy is that the spectra are pressure and temperature dependent. Increasing pressure will broaden ...
  51. [51]
    Sampling Techniques for FTIR Spectroscopy - JASCO Inc
    For liquids, the sample can be injected into a liquid cell, sandwiched between KBr salt plates, or applied to an IR transparent card. For gases, the sample can ...
  52. [52]
    FTIR: Transmission vs ATR spectroscopy | Animated Guides
    NaCl or CaF2 windows are typically used when liquids are analyzed through transmission. However, when using the ATR technique, the sample preparation is usually ...Missing: broadening | Show results with:broadening
  53. [53]
    [PDF] KBr Pellet Method - Shimadzu
    The KBr pellet method is mainly used to measure solid samples. Pellet methods utilize the pellet's property of becoming clear to infrared light when pressure is ...Missing: DRIFTS pyrolysis polymers
  54. [54]
    Guide to FT-IR Spectroscopy | Bruker
    DRIFTS requires careful sample preparation, so it is more difficult to perform, but it yields excellent quantitative results when analyzing solid samples such ...Missing: pyrolysis | Show results with:pyrolysis
  55. [55]
    By Infrared Spectra of Their Pyrolysis Products | Analytical Chemistry
    A simple microtechnique for obtaining pyrolyzate infrared spectra from polymeric materials. Applied Spectroscopy 1990, 44 (9) , 1587-1588.
  56. [56]
    Common Problems and Precautions in the Operation of FTIR ...
    Here we sorts out the common problems in sample preparation of FTIR ... Water vapor bands: These appear as sharp peaks in the regions around 3700-3500 ...Missing: OH subtraction
  57. [57]
    A Rapid and Simplified Approach to Correct Atmospheric ...
    Oct 31, 2024 · IR absorption bands due to water vapor in the mid-IR regions often obscure important spectral features of the sample. Here, we provide a ...
  58. [58]
    [PDF] Analysis of Polymers by ATR/FT-IR Spectroscopy - PIKE Technologies
    powerful sampling tool for the analysis of polymers. ATR eliminates sample preparation generally required for FT-IR analysis by transmission sampling ...Missing: pyrolysis | Show results with:pyrolysis
  59. [59]
    Applications of ATR-FTIR spectroscopic imaging to biomedical ...
    One of the key advantages of ATR-FTIR imaging is that it requires minimal or no sample preparation prior to spectral measurements. This is due to the fact that ...2. Ftir Imaging For The... · 3. Diamond Atr-Ftir Imaging · AcknowledgmentsMissing: post- | Show results with:post-<|control11|><|separator|>
  60. [60]
    Exploring the Steps of Infrared (IR) Spectral Analysis - PubMed Central
    Sep 30, 2023 · IR spectroscopy measures how molecules interact with infrared light, providing a measurement of the vibrational states of the molecules.
  61. [61]
    High-definition Fourier Transform Infrared (FT-IR) Spectroscopic ...
    Jan 21, 2015 · Go to the 'Electronics' tab and select an appropriate spectral resolution, typically of 4 cm-1 or 8 cm-1 for tissue. Go to the 'Background' tab ...
  62. [62]
    10.2: Improving the Signal-to-Noise Ratio - Chemistry LibreTexts
    Sep 12, 2021 · Signal averaging works well when the time it takes to collect a single scan is short and when the analyte's signal is stable with respect to ...
  63. [63]
    [PDF] How experimental decisions affect the signal-to-noise ratio
    Moreover, it is important to recognize the trade-offs between SNR and user-settable parameters, including number of scans, spectral resolution, and aperture.
  64. [64]
    Why and How Savitzky–Golay Filters Should Be Replaced
    Feb 17, 2022 · Savitzky–Golay (SG) filtering, based on local least-squares fitting of the data by polynomials, is a popular method for smoothing data and calculations of ...Introduction · Filters That Can Replace... · Applications · Supporting Information
  65. [65]
    Algorithms Used for Data Processing in FTIR - Shimadzu
    In this article, I will present some of the algorithms used for a variety of data processing operations applied to spectra in FTIR.
  66. [66]
    [PDF] beer's law limitation - harvey.pdf
    Cali- bration curves based on Beer's law are common in quantitative analyses. 10B.4 Beer's Law and Multicomponent Samples. We can extend Beer's law to a ...
  67. [67]
    FT-IR Search Algorithm – Assessing the Quality of a Match
    Various algorithms can be used to create a hit quality index (HQI), which is a measure of how well the query spectrum compares against each reference spectrum.Missing: qualitative picking
  68. [68]
    Evaluated Infrared Spectra - the NIST WebBook
    Evaluated Infrared Reference Spectra. S.E. Stein. This collection of infrared spectra was originally edited and published by the Coblentz Society.Missing: qualitative picking
  69. [69]
    Quantitative Mineral Analysis by FTIR Spectroscopy
    All multicomponent analyses are based on Beer's law, and the absorbance at a specific wavenumber is the sum of the absorbance of all sample components that ...<|separator|>
  70. [70]
    Multivariate Least-Squares Methods Applied to the Quantitative ...
    A linear least-squares approximation to nonlinearities in the Beer-Lambert law is made by allowing the reference spectra to be a set of known mixtures.
  71. [71]
    Application of New Least-squares Methods for the Quantitative ...
    By fitting only regions of the spectra that follow Beer's Law, accurate results can be obtained using three of the fitting methods even when baselines are not ...
  72. [72]
    A Process for Successful Infrared Spectral Interpretation
    Jan 1, 2016 · Spectral artifacts are peaks not caused by the sample. The most common types of spectral artifacts in IR spectra are water vapor and CO 2 ...Missing: removal stray light
  73. [73]
    Stray Light Correction in the Optical Spectroscopy of Crystals - NIH
    In this paper, we describe a method that allows use of the entire crystal by correcting the distorted spectrum.Missing: CO2 H2O
  74. [74]
    Far-Infrared Spectroscopy - an overview | ScienceDirect Topics
    Among these, the mid-IR is the most widely employed region for the monitoring of parameters of fats and oils as well as the most used technique for ...
  75. [75]
    Near-Infrared Spectroscopy Guide & FTIR Tips | Shimadzu
    Characteristics of Near-Infrared Light Absorption​​ However, near-infrared light absorption is much weaker in intensity as compared with mid-infrared light ...
  76. [76]
    Near Infrared - an overview | ScienceDirect Topics
    The near infrared region is populated with overtone and combination bands of lower frequency fundamentals below 4,000 cm−1. These bands are considerably weaker ...
  77. [77]
    [PDF] Exploring Dynamics in the Far-Infrared with Terahertz Spectroscopy
    Terahertz (THz) spectroscopy, also known as far-infrared, covers 0.3 to 20 THz, and is used to explore dynamics in liquids and semiconductors. It is a tabletop ...Missing: ligand | Show results with:ligand<|control11|><|separator|>
  78. [78]
    Lineshapes in IR and Raman Spectroscopy: A Primer
    Nov 1, 2015 · The band also broadens because of the rapid energy relaxation caused by the hydrogen bonds (shorter lifetime = broader peak).Missing: rotational | Show results with:rotational
  79. [79]
    IR Absorption Table
    Table of IR Absorptions. Functional Group, Characteristic Absorption(s) (cm-1), Notes. Alkyl C-H Stretch, 2950 - 2850 (m or s), Alkane C-H bonds are fairly ...
  80. [80]
    IR Absorption Frequencies - NIU - Department of Chemistry and ...
    Typical IR Absorption Frequencies For Common Functional Groups ; Ester, 1750–1730 (s) ; Amide, 1670–1640 (s) ; Anhydride, 1810 and 1760 (s) ; Acid Chloride, 1800 (s).
  81. [81]
    [PDF] Transition Metal Carbonyls
    The CO stretching frequency in the IR spectrum falls to 1720–1850 cm−1 on bridging. Type of CO v(CO) cm-1. Free CO. 2143 terminal M-CO.
  82. [82]
    The Urey — Bradley Force Field: Its Significance and Application
    The Urey-Bradley force (UBF) field which consists of stretching and bending force constants, as well as repulsive force constants between nonbonded atoms.
  83. [83]
    Re‐evaluation of the bond length–bond strength rule: The stronger ...
    Oct 29, 2015 · ... force constant k and bond length r exists. This so-called Badger Rule1, 2 was the basis for the tenet relating shorter bonds to stronger bonds.
  84. [84]
    Isotopes in assigning Vibrational Spectra - Oregon State University
    Hooke's Law: the vibrational frequency is equal to 1 over 2 pi c times the square root of the force constant ober the reduced mass.
  85. [85]
    Revisiting Isotope Effects in Infrared Spectra of H2O, HDO and D2O
    Sep 29, 2025 · Isotope effects are widely used in infrared analyses of a range of chemical systems. Water molecules with three active modes are one of the ...Introduction · Figure 3 · Basic Theoretical...
  86. [86]
    Two-Dimensional Infrared Spectra of the 13 C 18 O Isotopomers of ...
    The amide-I mode is largely a carbonyl stretch, and so its frequency is significantly modified by replacement of 12C by 13C (shift −38 cm-1), 16O by 18O (shift ...
  87. [87]
    Isotope-edited IR spectroscopy for the study of membrane proteins
    Isotope editing, through its ability to resolve individual vibrations, establishes FTIR as a method that is capable of yielding accurate structural data on ...Secondary Structure · Structure And Orientation · 2d Ir Spectroscopy
  88. [88]
    Forensic Applications of Isotope Ratio Mass Spectrometry
    May 4, 2016 · Forensic investigators have used IRMS to measure a variety of materials, such as drugs, explosives, food, and human remains.Missing: codes | Show results with:codes
  89. [89]
    Using 2D-IR Spectroscopy to Measure the Structure, Dynamics, and ...
    In this tutorial review, the method and development of ultrafast 2D-IR spectroscopy is discussed, including an introduction to the two main exptl.
  90. [90]
    Coherent 2D IR Spectroscopy: Molecular Structure and Dynamics in ...
    Two-dimensional infrared (2D IR) vibrational spectroscopy is an experimental tool for investigating molecular dynamics in solution on a picosecond time scale.
  91. [91]
    Dual-frequency 2D-IR spectroscopy heterodyned photon ... - PNAS
    In this article we report a dual-frequency 2D-IR spectroscopy heterodyned photon-echo experiment. The 2D-IR spectra, which require manipulation of the ...
  92. [92]
    Applications of two-dimensional infrared spectroscopy - Analyst ...
    In this minireview, we briefly outline the fundamental principles and experimental procedures of 2D IR spectroscopy. Using illustrative example studies, we ...Missing: paper | Show results with:paper
  93. [93]
    Applications of Two-Dimensional Infrared Spectroscopy - PMC
    This minireview provides a brief outline of the general principles, experimental implementation, and information available from two-dimensional (2D) IR ...
  94. [94]
    Two-dimensional infrared-Raman spectroscopy as a probe ... - Nature
    Apr 7, 2023 · We identify a relationship between the tetrahedral order of liquid water and its two-dimensional IR-IR-Raman (IIR) spectrum.Missing: seminal | Show results with:seminal<|control11|><|separator|>
  95. [95]
    Sadtler Database of Infrared Spectra. FTIR Spectral Library
    Today the Sadtler collection of infrared spectra contains over 225.000 spectra and is being extended. The most popular product of Sadtler is the HaveItAll(R) ...
  96. [96]
    Library Searching - Spectroscopy Online
    Mar 1, 2021 · The company with perhaps the largest collection of infrared spectral libraries is what was Sadtler, became Bio-Rad, and is now part of Wiley (3– ...
  97. [97]
    What Is Infrared Spectroscopy? Fundamentals & Applications - Excedr
    Jan 31, 2024 · The IR spectrum helps identify functional groups and impurities and provides quantitative data, enabling concentration measurements of specific ...
  98. [98]
    Infrared Spectroscopy: Principles and Applications in Organic ...
    Dec 28, 2024 · Infrared spectroscopy is used to determine chemical structures, monitor reaction progress, detect impurities, evaluate polymer ...
  99. [99]
    IR-EcoSpectra: Exploring sustainable ex situ and in situ FTIR ... - NIH
    FTIR helps reduce the undesired derivatives by ensuring reactions proceed efficiently and by monitoring the purity of products.
  100. [100]
    How is GC-IR Used? - Chromatography Today
    Feb 25, 2022 · GC-IR stands for gas chromatography infrared spectroscopy. It combines gas chromatography, as a separation technique, with characterisation power of infrared ...
  101. [101]
    Introduction to hyphenated techniques and their applications ... - NIH
    The hyphenated technique developed from the coupling of an LC and the detection method infrared spectrometry (IR) or FTIR is known as LC-IR or HPLC-IR. While ...
  102. [102]
    GPC–IR Hyphenated Technology to Characterize Copolymers and ...
    This application case (4) shows how the GPC–IR hyphenated technique is used to deformulate complex polymer mixtures for competitive analysis, IP protection, ...
  103. [103]
    Understanding Infrared and Raman Spectra of Pharmaceutical ...
    Mar 1, 2011 · X-ray diffraction and spectroscopic techniques are used to identify pure forms and mixtures of polymorphs.
  104. [104]
    Infrared and Raman spectroscopy techniques applied to ...
    This review summarizes the recent trends and developments of infrared and Raman spectroscopy applied to the identification of explosives
  105. [105]
    Analysis of Explosives Using Infrared Spectroscopy - ACS Publications
    Remote mid IR Photoacoustic Spectroscopy for the detection of explosive ... spectroscopy in the forensic identification of illicit drugs and explosives.
  106. [106]
    Infrared Spectroscopy of Polymers XII: Polyaramids and Slip Agents
    May 1, 2023 · In this column, we examine the spectra of polyaramids, which are polyamides that contain an aromatic ring in their chain.
  107. [107]
    [PDF] Combined Spectroscopy Problems With Solutions
    By understanding common problems such as spectral interference, calibration issues, and data integration difficulties, scientists can implement targeted ...
  108. [108]
    [PDF] Chapter 7. Fourier Transform Infrared Spectroscopy - VTechWorks
    Our approach was based on the examination of the ratios of the integrated intensities of the doublet bands at 730 and 720 cm-1. This ratio, I730/I720, for the ...
  109. [109]
  110. [110]
    Infrared Spectroscopy in Analysis of Polymer Degradation - Edge
    Sep 15, 2006 · Infrared (IR) spectroscopy locates the spectral position (frequency) and intensity of absorbed (or emitted) radiation arising from molecular ...
  111. [111]
    Reflection-Absorption Infrared Spectroscopy - ScienceDirect.com
    Infrared reflection absorption spectroscopy (IRAS) is defined as a technique used to investigate molecular adsorbates on metal surfaces.<|separator|>
  112. [112]
    Application of polarization-modulated Fourier transform infrared ...
    The utility of polarization modulation infrared reflection absorption spectroscopy (PM-IRRAS) in surface and in situ studies: new data processing and ...
  113. [113]
    Advances of surface-enhanced Raman and IR spectroscopies - Nature
    Aug 4, 2021 · Raman and infrared (IR) spectroscopy are powerful analytical techniques, but have intrinsically low detection sensitivity.
  114. [114]
    Confinement effects on the vibrational properties of III-V and II-VI ...
    We describe the size dependence of the vibrations and find good agreement for Raman shifts and for the frequency of coherent acoustic modes with experiments.
  115. [115]
    Orientation analysis of polymer thin films on metal surfaces via IR ...
    Aug 30, 2022 · We propose infrared (IR) absorbance-spectroscopy near metal surfaces to distinguish the orientation of different molecular segments.Missing: nanomaterials | Show results with:nanomaterials
  116. [116]
    Influence of Si doping level on the Raman and IR reflectivity spectra ...
    Quite generally, the doping atom itself disturbs the perfection of the semiconductor lattice and ultimately at high concentration generates defects. On the ...
  117. [117]
    Infrared spectroscopy and microscopy in cancer research and ... - NIH
    This review is aimed at illustrating some principles of Fourier transform (FT) IR spectroscopy and microscopy (microFT-IR) to interrogate molecules in specimen ...Missing: 1980s | Show results with:1980s
  118. [118]
    Fourier Transform Infrared Spectroscopy as a Cancer Screening and ...
    Jan 1, 2020 · Furthermore, FTIR spectroscopy has been utilized for cancer diagnosis by analyzing other types of biological materials besides the commonly ...
  119. [119]
    Fourier Transform Infrared Microscopy Enables Guidance of ... - Nature
    Jan 10, 2018 · We show that FTIR microscopy can automatically guide high-resolution MSI data acquisition and interpretation without requiring prior histopathological tissue ...
  120. [120]
    Noninvasive Blood Glucose Monitoring Systems Using Near ... - NIH
    This review paper briefly discusses the noninvasive glucose measuring technologies and their related research work.
  121. [121]
    Obtaining information about protein secondary structures in ... - Nature
    Feb 5, 2015 · In this protocol, we have detailed the principles that underlie the determination of protein secondary structure by FTIR spectroscopy.
  122. [122]
    FT-IR Open-Path Monitoring Guidance Document: Third Edition
    Validation of a Method for Estimating Pollution Emission Rates Using Open-Path FTIR Spectroscopy and Modeling Techniques. ... Long-Path Infrared Spectroscopy for ...Missing: NOx | Show results with:NOx
  123. [123]
    Greenhouse Gas Mapping - AVIRIS-Next Generation
    May 22, 2024 · AVIRIS-NG and AVIRIS-C can be used to measure increased concentrations of atmospheric methane and carbon dioxide.
  124. [124]
    Water quality analysis of organic pollutants in electronic industry ...
    Infrared and NMR analyses suggested that the wastewater contained diverse pollutants, including phenols, esters, aromatic hydrocarbons, aliphatic hydrocarbons, ...
  125. [125]
    Near-Infrared Spectroscopy in Bio-Applications - PMC - NIH
    NIR spectroscopy offers immense potential in various bio-applications with its primary advantages; rapid analysis, wide-applicability to variety of samples.
  126. [126]
    Using Fourier transform IR spectroscopy to analyze biological ...
    The main steps required to analyze a sample of interest are sample preparation, instrumental setting, acquisition of spectra and data processing (Fig. 4).
  127. [127]
    Computational Thermochemistry: Scale Factor Databases and Scale ...
    Aug 20, 2010 · Mid-IR spectra predicted using LSDA, BLYP, and B3LYP force fields are of significantly different quality, the B3LYP force field yielding ...Introduction · Methodology · Results and Discussion · Applications of Universal...
  128. [128]
    Combining the Double-Harmonic Approximation with Machine ...
    Dec 12, 2024 · Analysis tools for the combination of computed Hessian matrices and dipole derivatives into IR spectra can be found at https://github.com/pprcht ...
  129. [129]
    The ORCA quantum chemistry program package - AIP Publishing
    Jun 12, 2020 · ORCA is a widely used program in various areas of chemistry and spectroscopy with a current user base of over 22 000 registered users in academic research and ...
  130. [130]
    Influence of the computational methods in the conformational ...
    Oct 28, 2025 · Recent advancements in experimental techniques have provided spectra for single conformers for relatively large flexible molecules, ...Missing: virtual | Show results with:virtual
  131. [131]
    Theoretical Infrared Spectra: Quantitative Similarity Measures and ...
    We present here the calculation of the infrared spectra of a database of 703 gas phase compounds with four different force fields (CGenFF, GAFF-BCC, GAFF-ESP, ...Missing: screening | Show results with:screening
  132. [132]
    Adaptive vibrational configuration interaction (A-VCI) - NASA ADS
    A new variational algorithm called adaptive vibrational configuration interaction (A-VCI) intended for the resolution of the vibrational Schrödinger ...Missing: spectroscopy | Show results with:spectroscopy
  133. [133]
  134. [134]
    Product Review: Portable FTIR spectrometers get moving
    Design considerations for portable mid‐infrared FTIR spectrometers used for in‐field identifications of threat materials.Missing: seminal | Show results with:seminal
  135. [135]
    ProtectIR | Handheld FTIR for Solid and Liquid Identification
    Designed for field use, ProtectIR enables hazmat response teams, military, and first responders to analyze bulk substances in seconds. Wistia video ...
  136. [136]
    Infrared Spectroscopic Imaging: The Next Generation - PMC
    Since FPA detectors are two dimensional, it also is possible to spectroscopically and spatially map the sample at the same time. A hybrid imaging approach has ...
  137. [137]
    Progress in focal plane array technologies - ScienceDirect.com
    This paper presents progress in optical detector technology of focal plane arrays during the past twenty years.
  138. [138]
    High-resolution Fourier-transform infrared chemical imaging ... - NIH
    We present an FTIR imaging approach that substantially extends current capabilities by combining multiple synchrotron beams with wide-field detection.
  139. [139]
    A bright future for synchrotron imaging | Nature Photonics
    A synchrotron source has the capability of providing infrared light through a 10-μm pinhole that is 2–3 orders of magnitude brighter than a conventional Globar.
  140. [140]
    Guide to FT-IR Microscopy | Bruker
    Carefully arranged mirrors can even be used to create objectives, which are called Cassegrain objectives. Inside the infrared microscope.
  141. [141]
    [PDF] INFRARED MICROSCOPY - JASCO Inc
    Fig. 22 IQ Mapping allows the sample measurement area to be moved while using a stationary Cassegrain objective to build up the sample image.
  142. [142]
    Advances in ATR-FTIR Spectroscopic Imaging for the Analysis ... - NIH
    Jun 12, 2023 · Diamond, on the other hand, is a good choice in a wide pH range, withstanding high pressure and temperatures whilst also offering the highest ...Missing: post- | Show results with:post-
  143. [143]
    Development of ATR Objectives for Wide-Area Micro-FTIR Imaging ...
    Jun 10, 2020 · A Cassegrain objective is combined with a zinc sulfide or diamond ATR prism, which are transparent to light in both the visible and IR ...
  144. [144]
    Photoacoustic infrared spectroscopy of industrial materials (invited)
    Jan 1, 2003 · Five examples of the use of PA spectroscopy for the characterization of industrial materials are described, with particular emphasis on ...
  145. [145]
    Photoacoustic Spectroscopy - an overview | ScienceDirect Topics
    Photoacoustic spectroscopy (PAS) is based on the principle that modulated IR radiation striking the surface of a sample will cause the surface to alternately ...
  146. [146]
    High-throughput terahertz imaging: progress and challenges - Nature
    Sep 15, 2023 · Here, we review the development of terahertz imaging technologies from both hardware and computational imaging perspectives.
  147. [147]
    Recent advances in terahertz imaging: 1999 to 2021
    Dec 30, 2021 · We discuss the progress in the field of THz imaging based on time-domain spectroscopy during the last 20 years emphasizing several highlights.<|control11|><|separator|>
  148. [148]
    Drone-Mounted Infrared Camera Sees Invisible Methane Leaks in ...
    Jul 9, 2025 · Researchers in Scotland have developed a drone-mounted infrared imaging system that can detect and map methane gas leaks in real time from up to 13.6 meters ...
  149. [149]
    Autonomous multi-species environmental gas sensing using drone ...
    Mar 19, 2019 · A drone uses a broadband Fourier-transform infrared spectrometer for rapid, multi-species gas sensing, achieving 18-ppm sensitivity and ...
  150. [150]
    Characterising the performance of a drone-mounted real-time ...
    Mar 13, 2025 · We demonstrate a methane gas imaging system mounted to an unmanned aerial vehicle (UAV) that is shown to perform real-time detection at distances up to 10m ...
  151. [151]
    A compact QCL spectrometer for mobile, high-precision methane ...
    Sep 7, 2020 · A compact and lightweight mid-infrared laser absorption spectrometer has been developed as a mobile sensing platform for high-precision atmospheric methane ...