Fact-checked by Grok 2 weeks ago

Fourier-transform spectroscopy

Fourier-transform spectroscopy (FTS) is a measurement technique whereby spectra are collected based on measurements of the of a radiative source using time- or space-domain data, followed by application of the mathematical to obtain the . In the optical domain, it obtains the , visible, or of a substance by measuring the interference pattern, or interferogram, produced when light passes through a two-beam interferometer, such as a , and then reconstructing the via the inverse . In this optical method, broadband radiation from a source is divided into two beams by a ; one beam travels to a fixed mirror and the other to a moving mirror, after which the beams recombine to form an interferogram that encodes all wavelengths simultaneously as a function of the path difference between the mirrors. The resulting interferogram is digitized and processed by a computer using the to yield the intensity as a function of or , providing a complete in a single measurement. The origins of FTS trace back to the late 19th century with Albert A. Michelson's development of the interferometer for measuring light wavelengths and visibility functions, laying the groundwork for interferometric spectroscopy. Significant advancements occurred in the mid-20th century: Peter Fellgett proposed the multiplex principle in 1951, recognizing that interferometers could measure multiple wavelengths at once to improve signal-to-noise ratio for weak sources, an idea known as the Fellgett advantage. Pierre Jacquinot further developed the concept in 1958 by emphasizing the high throughput or étendue of interferometers compared to dispersive slits, termed the Jacquinot advantage, which allows greater light collection efficiency. Janine and Pierre Connes advanced practical implementations in the 1960s, achieving high-resolution planetary spectra and demonstrating wavelength accuracy through internal laser referencing, known as the Connes advantage; their work was pivotal in transitioning FTS from theoretical to routine use. The 1965 introduction of the Cooley-Tukey fast Fourier transform algorithm revolutionized computational feasibility, enabling rapid processing of interferograms and widespread adoption of FTS. Compared to traditional dispersive using prisms or gratings, optical FTS offers several key , including faster acquisition times due to simultaneous measurement of all frequencies (multiplex ), higher for low-light applications, superior resolution limited only by the maximum path difference in the interferometer, and reduced . These benefits from the interferometer's to use large apertures without , achieving up to 200 times greater light-gathering power in the region, and its inherent photometric accuracy. In Fourier-transform (FTIR) , a common variant, these features have made it the preferred method for routine analysis, scanning full spectra in seconds rather than minutes. FTS finds broad applications across scientific fields, including molecular identification through or spectra in and , where it serves as a "fingerprint" for compounds by quantifying components in mixtures and assessing sample quality. FTS also encompasses time-domain methods in (NMR) and other pulsed spectroscopies, enabling high-resolution studies in and physics. In atmospheric and environmental research, it analyzes trace gases like and pollutants with high precision. Astronomical observations benefit from its sensitivity to weak sources, as demonstrated in early high-resolution studies of planetary atmospheres, such as Venus's CO₂ bands. Additionally, FTS extends to biological processes, production, and medical diagnostics via FTIR for non-invasive tissue analysis. Today, it remains a cornerstone technique due to its versatility, with ongoing developments in portable and high-resolution instruments enhancing its impact.

Introduction

Definition and Overview

Fourier-transform spectroscopy (FTS) is a measurement technique that records the interferogram—an interference pattern generated by combining light waves with varying path lengths—and applies the Fourier transform to convert this time- or path-domain signal into a frequency-domain spectrum, revealing the intensity distribution across different wavelengths or frequencies. This approach enables the simultaneous acquisition of data across the entire spectral range, offering advantages such as improved signal-to-noise ratio (known as the Fellgett or multiplex advantage) and enhanced light-gathering power compared to traditional dispersive methods that scan wavelengths sequentially. FTS is widely applied in fields like infrared, ultraviolet, visible, and microwave spectroscopy for analyzing molecular structures, atmospheric composition, and material properties. At its core, an FTS system consists of a light source, a that divides the incoming into two beams, a fixed mirror and a moving mirror to introduce a controllable difference, and a detector that captures the intensity of the recombined beams as the mirror translates. In a typical , light from the source strikes the beam splitter, which transmits one portion to the fixed mirror and reflects the other to the moving mirror; the beams then recombine at the splitter, pass through or interact with the sample, and reach the detector to form the interferogram, which is digitally processed to produce the spectrum. This configuration leverages principles, where constructive and destructive encodes spectral information. Prerequisite to understanding FTS is the distinction between spectral representations in (typically in nanometers or micrometers) and frequency (often expressed as in cm⁻¹, the reciprocal of ), with FTS inherently yielding results in the for direct interpretation of vibrational or transitions. The technique supports both modes, where spectra arise from light emitted by excited samples such as gases or plasmas, and modes, where transmitted light through a sample reveals absorbed frequencies corresponding to molecular bonds. Although the term "Fourier-transform spectroscopy" emerged in the amid advancements in computing that enabled practical , its foundational interferometric principles originated in 19th-century work by .

Historical Background

The foundations of Fourier-transform spectroscopy (FTS) were laid in the late through the work of , who developed the in the 1880s primarily to measure the of with high precision. Michelson's instrument, initially designed for the Michelson-Morley experiment to detect the luminiferous ether, demonstrated the interference patterns essential for later spectroscopic applications, marking the inception of interferometric techniques in optical measurements. Advancements in the mid-20th century transformed these early concepts into practical . In 1947, Marcel J. E. Golay invented the Golay cell, a sensitive pneumatic detector for radiation that overcame limitations of previous detectors and enabled the recording of interferograms in the region, facilitating the practical implementation of FTS. This was complemented by Peter Fellgett's 1951 PhD thesis at the , where he proposed the multiplex advantage of —allowing simultaneous measurement of all wavelengths to improve —and performed the first numerical Fourier transformation of an interferogram to derive a spectrum. In 1958, Pierre Jacquinot emphasized the high throughput or Jacquinot advantage of interferometers, which allows greater light collection efficiency compared to dispersive instruments with slits. The 1965 introduction of the Cooley-Tukey algorithm revolutionized the computational feasibility of processing interferograms. Janine and Pierre Connes advanced practical implementations in the 1960s, achieving high-resolution planetary spectra and demonstrating wavelength accuracy through internal laser referencing, known as the Connes advantage; their work was pivotal in transitioning FTS from theoretical to routine use. The 1960s saw the adoption of FTS in astronomy, driven by improvements in computing and instrumentation. Pioneering observations by and Janine Connes using a custom two-beam interferometer yielded the first high-resolution near-infrared FTS spectra of in 1966, revealing detailed absorption bands in its atmosphere and demonstrating the technique's potential for planetary studies. Concurrently, facilities like the began integrating FTS into solar telescopes in the early 1970s, with early systems supporting atmospheric and stellar . By the 1970s, the availability of minicomputers such as the PDP-8 enabled routine laboratory use of FTS, with commercial instruments like Digilab's FTS-14 (introduced in 1969) making the technology accessible beyond specialized observatories. Pulsed FTS emerged prominently in the 1970s through applications in (NMR), where advanced Fourier-transform methods to enhance sensitivity and resolution. Ernst's innovations, starting from his 1966 work at and expanding in the early 1970s, allowed rapid acquisition of time-domain signals (free induction decays) followed by Fourier transformation to yield frequency spectra, revolutionizing NMR and earning him the 1991 . Recent developments up to 2025 have focused on enhancing resolution and speed through integration with advanced sources and detectors. facilities have enabled synchrotron-based FTS (SR-FTIR) for high-brightness, broadband infrared microspectroscopy, achieving sub-micrometer in applications like biomolecular imaging. Quantum detectors, such as superconducting single-photon detectors, have improved sensitivity in the mid- and far-infrared, while post-2020 computational advances in algorithms—leveraging GPU acceleration—have enabled real-time spectrum processing for dynamic systems like transient molecular events.

Theoretical Foundations

Interferometry Basics

in Fourier-transform spectroscopy relies on the fundamental principles of , where waves from a common source superimpose to produce patterns of reinforcement or cancellation depending on their relative s. Constructive interference occurs when waves are in phase, resulting in maximum , while destructive interference happens when they are out of phase by π radians, leading to minimum . These effects arise from differences in optical lengths traveled by the waves, which introduce phase shifts proportional to the path δ. In a two-beam interferometer, the intensity of the recombined for a monochromatic source of λ is given by I(\delta) = I_0 \left[1 + \cos\left(\frac{2\pi \delta}{\lambda}\right)\right], where I_0 is the average and δ represents the difference between the beams. This equation describes how the pattern oscillates between maximum (constructive) and minimum (destructive) values as δ varies, with the cosine term capturing the phase-dependent . For a polychromatic source containing multiple wavelengths, the total forms a superposition of these individual patterns, encoding the spectral information into a single interferogram I(δ) = ∫ B(ν) [1 + cos(2π ν δ)] dν, where B(ν) is the as a function of ν = 1/λ. The interferogram's role in spectroscopy stems from this superposition, which multiplexes the contributions of all wavelengths into the path-difference domain, allowing the entire to be captured simultaneously. In mode, the interferogram directly reflects the distribution of emitted by the source, while in mode, the sample modulates the incident , imprinting its spectrum onto the resulting pattern. A key feature is the zero-path-difference point, where δ = 0 and the two beams have identical path lengths, yielding the maximum due to perfect constructive across all wavelengths and forming the central of the interferogram.

Mathematical Basis: Fourier Transform

The mathematical foundation of Fourier-transform spectroscopy relies on the relationship between the interferogram, which records as a of optical path difference δ, and the B(σ), representing as a of σ = 1/λ. For a source, the interferogram I(δ) is given by the I(\delta) = \int_0^\infty B(\sigma) \cos(2\pi \sigma \delta) \, d\sigma, assuming the is real and even, and neglecting the constant background term that arises from the zeroth-order . This equation describes how the interference pattern encodes the spectral information in the time- or path-difference domain, with the cosine term reflecting the difference for each component. To recover the spectrum from the measured interferogram, the inverse is applied: B(\sigma) = 2 \int_0^\infty I(\delta) \cos(2\pi \sigma \delta) \, d\delta, valid for even interferograms and under ideal conditions without . This pair establishes the interferogram as the forward cosine transform of the , enabling reconstruction through inversion. In practice, the integrals are approximated due to finite ranges and . For digital implementation, the continuous integrals are discretized into the discrete Fourier transform (DFT), where the interferogram is sampled at N points with interval Δδ, yielding a finite sequence I(n Δδ) for n = 0 to N-1. The spectrum is then computed as B(m \Delta\sigma) = \sum_{n=0}^{N-1} I(n \Delta\delta) \cos(2\pi m n / N), with Δσ = 1/(N Δδ), though the full complex form is often used for computational efficiency. The fast Fourier transform (FFT) algorithm, particularly the Cooley-Tukey radix-2 variant, efficiently computes the DFT in O(N log N) operations when N is a power of 2, revolutionizing digital processing in spectroscopy since its introduction in 1965 and subsequent application to interferogram analysis. Apodization is essential in digital processing to mitigate artifacts from the finite interferogram length, which otherwise produces Gibbs ringing (sidelobes) in the spectrum due to the sinc-like instrumental lineshape. A window function w(δ) is multiplied by I(δ) before transformation, such as the Boxcar (rectangular) window w(δ) = 1 for |δ| ≤ Δδ_max/2, which offers maximum resolution but high sidelobes (~22% of peak height), or the Hann window w(δ) = 0.5 [1 - cos(2π δ / Δδ_max)] for |δ| ≤ Δδ_max/2, which suppresses sidelobes to ~2.5% at the cost of ~50% resolution broadening. The choice balances ringing reduction against mainlobe widening, with the transform of the window convolving the true spectrum. Units in spectroscopy typically express path difference δ in cm and σ in cm⁻¹, so that the product σδ is dimensionless, aligning with the oscillatory argument in the cosine. Sampling must satisfy the Nyquist theorem, requiring interval Δδ ≤ 1/(2 σ_max) to avoid of the highest σ_max in the . The achievable Δσ, defined as the minimum resolvable separation (e.g., via ), is limited by the maximum path difference Δδ_max to Δσ ≈ 1 / Δδ_max for unapodized cases, where longer scans enhance but increase acquisition time and sensitivity.

Instrumentation

Continuous-Wave Michelson Spectrometer

The continuous-wave Michelson spectrometer forms the core instrumentation for traditional Fourier-transform spectroscopy (FTS), employing a broadband light source to generate an interferogram through interferometric modulation. The setup typically includes a collimated broadband source, such as an incandescent lamp for near-infrared wavelengths up to approximately 5 μm or a Nernst glower for mid- to far-infrared up to approximately 25 μm, which provides stable, continuous emission across the spectral range of interest. This radiation enters a 50/50 beam splitter, which divides the beam into two paths: one directed to a fixed mirror and the other to a moving mirror mounted on a precision translation stage. The reflected beams recombine at the splitter, producing interference patterns that are detected by a single-element photodetector, such as a mercury cadmium telluride (MCT) detector optimized for infrared sensitivity and rapid response. In operation, the moving mirror scans at a constant velocity along the optical path, systematically varying the path difference between the two arms and modulating the interference as a function of time, which is recorded as an interferogram representing detector signal versus path difference. This scanning ensures uniform sampling of the optical path difference (OPD), with velocity precisely controlled by a motorized drive to maintain consistent increments and avoid distortions in the interferogram. Beam splitters are selected based on the wavelength range; for mid-infrared applications, potassium bromide (KBr) substrates coated with germanium provide efficient 50/50 splitting up to 25 μm, though KBr's hygroscopic nature requires protective measures, and a compensator plate addresses dispersion-induced OPD asymmetry for balanced transmission and reflection. Typical scan lengths range from 1 to 10 cm of maximum OPD, enabling resolutions from 1 cm⁻¹ to 0.1 cm⁻¹, where finer resolution corresponds to longer scans to capture higher-frequency components in the interferogram. Calibration relies on a helium-neon (HeNe) laser reference beam, which generates evenly spaced fringes during the scan to track mirror position with sub-wavelength accuracy, correcting for any non-linearities and ensuring precise OPD sampling. This setup leverages basic interferometry principles to encode spectral information into the temporal domain.

Spectrum Acquisition and Processing

In Fourier-transform spectroscopy, spectrum acquisition begins with the digitization of the interferogram signal detected by the interferometer. The analog signal from the detector undergoes analog-to-digital conversion, typically triggered by the zero-crossings of a reference to ensure uniform sampling. A common sampling rate is one or two points per HeNe fringe (corresponding to OPD intervals of approximately 0.316 μm or 0.158 μm), which satisfies the and prevents of high-frequency components up to ~15800 cm⁻¹ or higher for the mid-infrared range. Zero-filling is often applied during data collection by appending zeros to the interferogram array, typically doubling or quadrupling the data points to a , which facilitates and improves the apparent in the final spectrum without altering the true . Pre-processing of the digitized interferogram addresses imperfections to ensure accurate . correction compensates for misalignment in the interferometer mirrors or detector response delays, which introduce phase errors manifesting as imaginary components in the complex ; this is achieved using methods like the Mertz correction, where a low-resolution phase spectrum from a central portion of the interferogram is applied multiplicatively to the full dataset. DC offset removal subtracts the mean value of the interferogram to eliminate baseline drifts caused by noise or detector , preventing artifacts in the spectral . These steps are typically implemented in dedicated software packages, such as those integrated with commercial FTIR systems from vendors like Thermo Fisher or , or general platforms like for custom setups. The core of spectrum processing involves applying the (FFT) to convert the time-domain interferogram into a frequency-domain , leveraging the mathematical to recover intensity as a function of . Interferograms are classified as double-sided (symmetric around the zero-path-difference point) or single-sided (recorded from zero path difference to maximum only); double-sided data naturally yield real-valued spectra with minimal phase errors, while single-sided interferograms require additional phase correction to avoid distortions, though they reduce acquisition time. FFT algorithms, such as the Cooley-Tukey method, efficiently handle these computations on arrays up to millions of points, enabling real-time processing in modern instruments. Common artifacts in processed spectra include ghosting from aperture reflections or multireflections in the sample, which appear as spurious peaks, and due to beyond the Nyquist (half the maximum sampling ). Atmospheric absorption lines from CO₂ (around 2350 cm⁻¹) and H₂O (broad bands near 3700–3600 cm⁻¹ and 1600 cm⁻¹) are subtracted by ratioing the sample spectrum against a background interferogram recorded under identical conditions, often automated in software to yield a clean spectrum. Apodization is applied prior to the FFT to mitigate spectral leakage from abrupt truncation of the interferogram, which causes ringing or sidelobes in sharp spectral features. The Happ-Genzel function, a trapezoidal apodization with continuous first derivatives, is widely used in infrared applications for its balance between suppressing sidelobes (to about 1–2% of peak height) and preserving resolution (slight broadening to 55–60% of the unapodized value). For absorption spectra, baseline correction follows the transform, using polynomial fitting or rubberband algorithms to remove sloping baselines from scattering or instrumental drift, ensuring accurate peak quantification. The final output is a spectrum in transmittance or absorbance units versus wavenumber (cm⁻¹). Wavenumber calibration relies on the precise HeNe laser reference, mapping path difference to spectral axis with accuracy better than 0.01 cm⁻¹; apparent absorbance, calculated as -log₁₀(T), may overestimate true values due to uncompensated reflections or scattering, necessitating corrections like Kubelka-Munk transformation for in diffuse reflectance modes.

Pulsed Fourier-Transform Spectroscopy

Operational Principles

In optical implementations of pulsed Fourier-transform spectroscopy (FTS), ultrashort pulses, typically in the to range, generated by mode-locked lasers such as Ti:sapphire oscillators, are employed rather than steady sources. Analogous principles apply in other domains, such as with radiofrequency pulses in the range. These pulses, often 50–100 fs in duration with repetition rates of 80–100 MHz, are split into a pump beam for excitation and a probe beam for detection, enabling the study of ultrafast transient phenomena. This pulsed operation facilitates direct measurement in the time domain, where the spectral content of the pulses allows simultaneous probing across a wide range of frequencies. The operational core involves time-domain interferometry, in which the short pulses excite transient coherences that produce a decaying signal, such as an waveform, which is captured over time scales of picoseconds. This signal is sampled at high , equivalent to terahertz rates (e.g., ~150 THz for 6.6 fs steps), using an optical delay line to incrementally vary the probe pulse timing relative to the signal. The resulting time-domain interferogram is then Fourier-transformed to yield the frequency-domain spectrum, leveraging the inverse relationship between time and frequency domains. Triggering and rely on precise alignment of the pulse train, with the repetition rate governing the coverage in advanced configurations like asynchronous optical sampling, where differential repetition rates (e.g., 1 GHz) define the and range up to the pulse bandwidth. Phase-stable referencing, often via common optical paths or electronic stabilization, preserves the of the measurement throughout the delay scan. The time of the transient signal inherently limits the , as longer decay times enable finer discrimination (Δν ≈ 1/τ_coherence). Detection employs time-resolved techniques, including photoconductive antennas that convert the THz or field into measurable photocurrents via femtosecond-gated carriers, or electro-optic sampling in nonlinear where the probe pulse is modulated by the signal . For higher speeds, streak cameras can capture the temporal evolution directly. is enhanced by averaging multiple pulse train acquisitions, typically thousands of scans, often with lock-in amplification to isolate the coherent signal from noise.

Free Induction Decay

In pulsed Fourier-transform spectroscopy, the free induction decay (FID) is the transient time-domain signal that arises after an excitation pulse, capturing the relaxation of coherences or transverse magnetization in the system. This signal encodes the frequency components of the through the of excited states at their characteristic frequencies, modulated by processes. The FID is mathematically described as I(t) = \sum_k A_k \exp(i \omega_k t) \exp(-t / T_2), where A_k represents the of the k-th component, \omega_k is its , and T_2 denotes the transverse relaxation time governing the . Applying the to this FID produces the frequency-domain , featuring lines with lineshapes whose is approximately $1 / (\pi T_2). In ultrafast Fourier-transform (FTIR) spectroscopy, the FID emerges from vibrational coherences induced by a pump-probe , where short pulses excite molecular vibrations, leading to a decaying that reveals dynamics. Similarly, in (NMR) spectroscopy, a radiofrequency pulse perturbs the magnetization into the , generating an FID that encodes chemical shifts and scalar couplings through the phase evolution of individual isochromats. To acquire the FID comprehensively, detection is employed, utilizing two phase-shifted receiver channels to simultaneously record real and imaginary components, thereby preserving information and avoiding folding artifacts. From the acquired FID, the transverse relaxation time T_2 can be determined by fitting the signal envelope to an model, providing insights into molecular mobility and interactions. Processing the FID prior to transformation often involves with functions like the shifted sine-bell, which tapers the signal to reduce from truncation while optimizing resolution for the inherent , particularly beneficial in noisy or limited-duration acquisitions.

Specific Examples

Fourier-transform (FT-NMR) exemplifies pulsed FTS in the radio-frequency domain, where a 90° radiofrequency pulse is applied to tip the net into the , generating a (FID) signal that is acquired over 1-100 ms to cover spectral ranges from 1-800 MHz depending on the strength. This setup requires precise shimming to achieve homogeneity, minimizing linewidth broadening and enabling high-resolution spectra of molecular structures. In two-dimensional FT spectroscopy, coherent multidimensional pulses create correlation spectra that reveal internuclear connectivities and dynamics, with techniques like photon echo methods developed in the enhancing resolution for complex systems such as biomolecules or excitons. These implementations use sequences of phase-coherent pulses to map couplings in two frequency dimensions, providing insights into energy transfer processes that one-dimensional methods cannot resolve. Ultrafast Fourier-transform infrared (FTIR) employs mid-infrared pulses generated from optical parametric oscillators (OPOs) to probe vibrational dynamics with 100 fs temporal resolution, capturing transient molecular responses in condensed phases. This approach allows real-time observation of bond breaking and changes in photochemical reactions, leveraging the nature of OPO outputs for simultaneous multi-mode tracking. A distinctive application involves time-resolved FTS for plasma diagnostics in laser-induced breakdown spectroscopy (LIBS), where mid-infrared dual-comb spectrometers acquire broadband spectra of species like methane and ethane in laser-generated plasmas during the 2020s. These setups enable quantitative analysis of transient emission and absorption lines, aiding in the characterization of reaction kinetics and composition in high-temperature environments. Addressing data challenges in pulsed FTS, oversampling the FID at rates exceeding the Nyquist limit captures rapid initial decays without truncation artifacts, preserving high-frequency components in the transformed spectrum. Noise reduction is further achieved through phase cycling of excitation pulses, which suppresses coherent artifacts and solvent signals, improving signal-to-noise ratios in multidimensional acquisitions.

Alternative Configurations

Stationary Interferometers

Stationary interferometers in Fourier-transform spectroscopy (FTS) employ fixed optical components to generate interferograms through spatial rather than temporal modulation of the difference. Unlike scanning designs, these systems produce interference fringes across a spatial , where the position on a detector array corresponds to varying path lengths, enabling the interferogram to be captured in a single exposure without mechanical motion. This approach relies on elements such as wedged s or tilted plates to create a linear variation in path difference, forming spatial fringes that encode the spectral information. For instance, a wedged introduces a gradual thickness change, resulting in path differences proportional to the lateral position, which are then "scanned" by the pixels of a detector array or by relative movement of the source. A notable is the birefringent wedge interferometer, where a tilted birefringent plate exploits the difference in refractive indices for and rays to produce the required spatial path variation. In this setup, the input beam is split into two orthogonally polarized components that propagate with a position-dependent delay, yielding an interferogram along one of the detector. This eliminates the need for a traditional and moving mirrors, simplifying construction while maintaining the core FTS principle of inversion to recover the . Early implementations in the explored designs for compact, stable setups, though modern variants prioritize spatial encoding for broader applicability. Spatial FTS with focal plane arrays extends this concept to imaging applications, particularly , where the entire interferogram is recorded simultaneously across a two-dimensional detector. Here, the spatial position directly maps to path difference, allowing reconstruction of spectral cubes that combine spatial and information without scanning. Detector arrays, such as or focal plane arrays in the , sample the fringes at high , enabling spectral recovery; for example, channel-dispersed designs use fixed prisms or gratings to separate wavelengths into parallel interferograms, boosting efficiency and resolution. These systems achieve resolving powers up to 1600 in the near-, corresponding to spectral resolutions of several cm⁻¹ (e.g., ~8 cm⁻¹ at 800 nm). The absence of in interferometers provides inherent insensitivity to , making them robust for environments where mechanical stability is challenging, such as deployments. is fundamentally set by the maximum path difference across the sampled , which for wedged or tilted elements scales with the wedge angle or tilt; a 1° tilt over a typical 1 cm can yield fringes supporting resolutions of tens of cm⁻¹ (e.g., ~30–50 cm⁻¹) in optimized setups by maximizing the number of interference cycles. This vibration resistance and compactness facilitate portable devices for on-site analysis, including handheld units for and , where rapid, single-shot acquisitions are essential.

Other Interferometric Designs

In addition to the conventional Michelson design, several alternative interferometric configurations have been developed for Fourier-transform spectroscopy (FTS), offering enhanced stability, compactness, or suitability for specific regimes through stationary or hybrid approaches. The Sagnac interferometer represents a circular common-path design that achieves path differences by splitting and recombining counter-propagating beams in a loop configuration, often modified for FTS by replacing one mirror with a transmission grating to generate tilts via . This setup produces Fizeau fringes that are imaged onto a detector, with the recovered via of the interferogram; heterodyning around a selected enables high without moving parts in static variants. Such systems provide exceptional insensitivity due to the common-path and have been applied in fiber-optic FTS for wavelengths around 1550 nm, where tunable gratings allow self-calibration over broad ranges. Mach-Zehnder-based interferometers adapt the dual-beam splitting architecture into a modified common-path form, where light is divided into parallel paths using beam splitters and mirrors, then recombined to form a spatial interferogram detected by an array sensor. This rejects common-mode vibrations effectively, as perturbations affect both arms equally, enabling robust operation in unstable environments without mechanical scanning. Applications in visible and UV benefit from the design's high throughput and adaptability to array detectors, achieving resolutions suitable for high-speed spectral imaging in the 400–700 nm range. Fabry-Pérot etalons operated in FT mode leverage multiple reflections between two parallel high-reflectivity mirrors to achieve high , enhancing contrast and effective . The etalon is quasi-stationary, with path differences introduced by piezo-tuning one mirror over distances up to 60 μm, generating an interferogram from the modulated output intensity that is Fourier-transformed to yield the . This yields resolutions around 2 nm at 532 nm, limited by the scan range, and supports compact imaging spectrometers with high luminosity for visible to near-IR applications. Lamellar grating interferometers, featuring movable parallel plates or s that divide the into multiple segments, extend FTS to challenging regimes like the UV by using all-mirror designs to minimize and . These wavefront-division systems produce interferograms through adjustable overlaps of beam segments, enabling operation from UV (down to ~200 nm) to far-IR with high étendue preservation. In UV applications, they facilitate high-resolution of solids and gases, with knife-edge prisms allowing split-ratio tuning for optimized signal-to-noise. Birefringent FTS configurations, often based on Savart or Wollaston prisms, introduce path differences via polarization-dependent refractive index variations, making them ideal for simultaneous spectropolarimetry. High-order retarders encode polarization states into channeled interferograms, recovered via to yield both spectral and polarimetric data in a single snapshot. In astronomical instruments, such as compact imaging spectropolarimeters, these designs enabled wavelength-dependent Stokes parameter measurements for studying in and , achieving resolutions of ~10 nm in the visible with robustness to alignment errors. These designs trade off throughput against complexity: while common-path and birefringent variants enhance and étendue (often >50% utilization), they require precise control or custom gratings, increasing fabrication challenges compared to simple two-beam systems. in multiple-reflection schemes, such as Fabry-Pérot, scales as \Delta \lambda / \lambda = 1/(2N) where N is the effective number of reflections, balancing finesse gains against reduced from higher reflectivity.

Advantages and Limitations

Multiplex and Throughput Advantages

Fourier-transform spectroscopy (FTS) offers the Fellgett advantage, also known as the multiplex advantage, which arises because all wavelengths across the spectrum are detected simultaneously by the interferometer, encoding the entire spectral information into a single interferogram. This multiplexed measurement improves the (SNR) compared to dispersive spectrometers, where wavelengths are scanned sequentially. In the detector-noise-limited regime, the SNR gain is approximately \sqrt{M}, where M is the number of spectral channels, as the noise from the detector is shared across all channels while the signal is coherently reconstructed via the . More precisely, the SNR in FTS exceeds that of a scanning dispersive instrument by a factor of \sqrt{N/8}, with N representing the number of spectral elements. The Fellgett advantage applies primarily in detector-noise-limited or source-limited (signal photon-noise-limited) regimes, where the multiplexed detection efficiently utilizes the available signal without amplifying uncorrelated noise. However, it diminishes or reverses in background-limited cases, such as when or background photon noise dominates, as this noise is also multiplexed, leading to an SNR penalty of up to \sqrt{M}. For broadband , typical SNR improvements from the Fellgett advantage range from 10 to 30 times over dispersive methods, depending on the number of resolved elements (e.g., \sqrt{M} \approx 50 for scanning 1 to 20 \mum at moderate resolution). Complementing the Fellgett advantage is the Jacquinot advantage, or throughput advantage, which stems from the absence of entrance and exit slits in FTS, allowing a larger —the product of aperture area A and \Omega—to pass through the instrument without sacrificing . In dispersive spectrometers, slits restrict the etendue to maintain , limiting light collection. For a circular in FTS, the etendue is given by A \Omega = \pi (D/2)^2 \sin^2 [\theta](/page/Theta), where D is the and \theta is the half-angle of , enabling up to 100 times higher throughput than equivalent -based systems. This results in practical throughput ratios of 100 to 200 for FT-IR versus instruments of similar . The combined Fellgett and Jacquinot advantages yield substantial SNR gains in FTS, often 10 to 30 times higher in broadband measurements compared to dispersive techniques, particularly when detector noise predominates. These benefits were first predicted by Fellgett in his 1951 PhD thesis at the , where he introduced the multiplex principle and demonstrated the first numerically Fourier-transformed interferogram. The prediction was validated in astronomical applications during the 1960s, as interferometric techniques enabled high-resolution observations that outperformed traditional dispersive methods.

Resolution, Sensitivity, and Limitations

The in Fourier-transform spectroscopy (FTS) is fundamentally determined by the maximum difference, Δδ, over which the interferogram is recorded, with the Δσ expressed as Δσ = 1/Δδ, where Δσ is in wavenumbers (cm⁻¹). Achieving higher requires longer path differences, but practical limits arise from mechanical stability and scan duration. functions are commonly applied to the interferogram to suppress and ringing artifacts inherent to the sinc instrumental line shape, though this broadens the effective line width and trades off some for improved fidelity. Sensitivity in FTS depends on the detector's (NEP), which quantifies the , and the total scan time, as extended acquisitions enable noise averaging to enhance (SNR). FTS particularly excels in low-light environments by simultaneously detecting all wavelengths, leveraging its throughput advantage for superior collection efficiency compared to dispersive methods. However, long scans necessary for high are prone to 1/f noise from detector dark current and electronics, which degrades low-frequency components and limits overall sensitivity. Key limitations include sampling errors that cause if the interferogram sampling rate is less than twice the maximum σ_max, causing high- components (> σ_max) to fold back into the lower- range, leading to spectral distortions and peaks. instabilities, often due to optical misalignment or drifts during scanning, introduce errors in the , manifesting as asymmetric lineshapes or baseline tilts. High- spectra impose significant computational demands; for instance, a of 0.01 cm⁻¹ requires a 100 cm path difference, generating interferograms with millions of data points that necessitate efficient algorithms and substantial processing resources. In open-path FTS configurations, atmospheric from variable by and other gases complicates quantification, while non-common-path designs exhibit heightened sensitivity to mechanical vibrations, which induce fluctuations and reduce accuracy. Mitigations for these challenges include adaptive sampling strategies, such as nonuniform or approaches, to reduce volume while preserving resolution against . Post-2020 advances in AI-based denoising, employing unsupervised machine learning techniques like () and (), have effectively suppressed noise and errors in FTS , improving SNR without excessive computational overhead.

Applications

Mid-Infrared and Far-Infrared Spectroscopy

Fourier-transform mid-infrared (MIR) spectroscopy, operating in the 400–4000 cm⁻¹ range, is widely employed for analyzing vibrational spectra of molecular species in gas and liquid phases using specialized transmission cells. These cells, often constructed with potassium bromide (KBr) windows that transmit effectively in the MIR region, allow for the interrogation of samples such as organic compounds to identify characteristic functional groups like carbonyls (C=O around 1700 cm⁻¹) and hydroxyls (O-H around 3400 cm⁻¹). Gas cells with path lengths up to 10 cm enable detection of trace volatiles, while liquid cells with thin spacers (typically 0.025–0.1 mm) accommodate neat liquids or solutions without dilution. To mitigate interference from atmospheric water vapor, which absorbs strongly at 1600–1800 cm⁻¹ and 3700–3900 cm⁻¹, instruments are routinely purged with dry nitrogen or air, significantly reducing baseline noise and improving signal-to-noise ratios. In far-infrared (FIR) Fourier-transform spectroscopy, covering 10–400 cm⁻¹, the technique targets low-energy rotational transitions, particularly in gaseous samples near the boundary. windows, offering high transmission (>80%) from 16 to 2500 μm (~4 to 625 cm⁻¹) with minimal , are standard for FIR cells and beam splitters to avoid material-induced artifacts. detectors, such as superconducting transition-edge sensors cooled to 4.2 K, are essential for FIR due to their sensitivity to weak thermal emissions and broad spectral response, achieving a of 1.2 × 10^{-13} W/√Hz. Purging remains critical here as well, since rotations overlap with sample signals around 100–200 cm⁻¹, potentially obscuring transitions in molecules like (PH₃) at 267 GHz (8.9 cm⁻¹). MIR-FTIR has been integral to since the 1990s, with U.S. Environmental Protection Agency (EPA) methods like TO-16 (open-path FTIR for fenceline pollutants) and (extractive FTIR for vapor-phase organics and inorganics such as and HCl) enabling real-time detection of hazardous air pollutants at parts-per-billion levels. In pharmaceutical and forensic analyses, FTIR adoption surged in the , leveraging its non-destructive nature to confirm substances via fingerprint spectra in under 1 minute. By 2025, handheld MIR-FTS devices, such as the Agilent 4300 series weighing ~2 kg, have advanced on-site forensics by identifying narcotics and explosives directly at scenes without , integrating libraries of over 13,000 spectra.

Terahertz and Microwave Spectroscopy

Fourier-transform spectroscopy (FTS) in the (THz) regime, spanning 0.1 to 10 THz, employs sources such as backward-wave oscillators (BWOs) and photoconductive antennas to generate and detect THz for interferometric measurements. BWOs provide tunable, monochromatic coherent with high output power and stability, enabling phase-sensitive THz spectroscopy where the complex is measured to derive spectra. Photoconductive antennas, excited by lasers, produce ultrafast THz pulses whose time-domain waveforms are Fourier-transformed to yield frequency-dependent spectra, facilitating applications like non-destructive imaging in security screening where THz waves penetrate non-conductive materials without . These techniques exploit the non-ionizing nature of THz waves for safe, high-resolution imaging of concealed objects. In the microwave regime (GHz frequencies), (FTMW) utilizes horn antennas for efficient coupling of chirped microwave pulses to gaseous samples, often in supersonic jets generated by pulsed nozzles to cool molecules and reduce rotational congestion. FTMW captures time-domain (FID) signals from polarized molecular ensembles, which are Fourier-transformed to produce high-resolution rotational spectra, aiding the identification of molecules in simulations for astronomical searches. The broadband capability of chirped-pulse FTMW allows simultaneous detection of multiple across 7-18 GHz or wider bands, enhancing its utility in molecular structure determination. Fourier-transform nuclear magnetic resonance (FT-NMR) operates as a form of FTS in the radiofrequency () domain, where high-field superconducting magnets—reaching up to 30.5 T (1.3 GHz for ¹H) as of 2025—enhance by increasing dispersion. In FT-NMR, the FID from spin coherences is Fourier-transformed to generate one-dimensional spectra, which extend to multi-dimensional variants (e.g., , ) that resolve complex couplings and correlations essential for determining protein structures in solution or solid states. These multi-dimensional spectra map internuclear distances and dynamics, enabling high-throughput applications. A notable advancement is cavity-enhanced FTMW spectroscopy, which integrates cryogenic superradiant cavities to amplify weak signals from trace gases, achieving detection sensitivities down to parts-per-billion (ppb) levels for atmospheric monitoring and environmental analysis. This enhancement boosts the effective path length and , allowing ultra-sensitive detection of transient without extensive averaging. Challenges in THz FTS include the need for cryogenic detectors, such as bolometers, to achieve sufficient sensitivity due to low energies, though recent cryogen-free pyroelectric receivers are addressing speed and cooling limitations. In FT-NMR, relaxation—particularly T2 (transverse) relaxation—poses challenges by broadening lines and limiting acquisition times, necessitating optimized sequences to mitigate decoherence in high-field environments.

Astronomical and Remote Sensing Uses

In astronomy, Fourier-transform spectroscopy (FTS) enables high-resolution observations of faint celestial objects, leveraging its coverage and multiplex advantage to detect molecular signatures in atmospheres. FTS instruments, such as the developed for detection and of exoplanets, combine interferometric techniques with array detectors to achieve simultaneous spectral and , allowing the isolation of planetary signals from stellar glare through with molecular templates. The Fellgett advantage, which improves by a factor of approximately √M (where M is the number of spectral resolution elements) in detector-noise-limited regimes, is particularly critical for these faint sources, enabling efficient use of limited time on ground-based facilities. Hybrid echelle-FTS designs further enhance this capability by integrating dispersive elements for order separation with interferometric , as demonstrated in spectral interferometers that bridge the gap between traditional echelle spectrographs and full FTS for mid-infrared astronomical applications. For instance, the SITELLE instrument on the Canada-France-Hawaii Telescope employs an imaging FTS to produce datacubes with resolutions up to R=5000 over 300–900 nm, facilitating studies of resolved stellar populations and galactic dynamics through integral field spectroscopy. In , open-path FTS systems provide path-integrated measurements of atmospheric greenhouse gases, offering high precision for monitoring urban and regional emissions without sample collection. The Total Carbon Column Observing Network (TCCON), operational since 2004, utilizes ground-based FTS instruments viewing direct solar spectra in the near-infrared to retrieve column-averaged dry-air mole fractions of CO₂ and CH₄ with a of 0.02 cm⁻¹, serving as a global reference for validating data. These systems integrate with differential optical absorption spectroscopy (DOAS) techniques to enhance detection in complex environments, such as urban boundary layers, by combining broadband FTS with targeted absorption modeling. An example is the open-path FTS observatory in , which measures CO₂ and CH₄ along kilometer-scale paths with precisions of 0.2 and 5 ppb, respectively, aiding flux quantification in heterogeneous terrains. Satellite-based FTS has revolutionized global CO₂ profiling, with the Greenhouse gases Observing SATellite (GOSAT), launched in 2009, employing the Thermal and Near-infrared Sensor for carbon Observation (TANSO-FTS) to achieve column-averaged CO₂ accuracies of ~0.3% from sun-glint and observations. Spatial FTS variants enable hyperspectral imaging by generating datacubes where each pixel encodes a full interferogram, convertible to spectra; static imaging FTS designs on small satellites produce gigapixel-scale volumes for and atmospheric analysis, benefiting from the throughput advantage for broadband . Recent advancements, including for cloud contamination rejection in GOSAT-2 data processing (launched 2018), have improved retrieval biases to <0.5 ppm in partially cloudy scenes, enhancing the utility of these datasets for modeling.

References

  1. [1]
    [PDF] FOURIER TRANSFORM SPECTROSCOPY - Johns Hopkins APL
    A . Michelson is used in a method called Fourier transform spectroscopy. The interferometer processes radiant energy into a Fourier transform of the spectrum.
  2. [2]
    [PDF] 19670017119.pdf - NASA Technical Reports Server
    The introduction of the Cooley-Tukey method is a significant land mark in the development of Fourier spectroscopy and should lead to its more rapid future ...Missing: key | Show results with:key
  3. [3]
    None
    ### Summary of FTIR Spectroscopy from the Document
  4. [4]
    Fourier Transform Infrared Spectroscopy and Microscopy
    Fourier Transform Infrared (FTIR) spectrometry is a valuable instrument for understanding atmospheric chemistry, biological processes, and bioenergy production.Missing: definition | Show results with:definition
  5. [5]
    [PDF] FOURIER TRANSFORM SPECTROSCOPY - Digital WPI
    Apr 24, 2019 · Today, Fourier transform spectroscopy has cemented itself as a widespread method of spectral collection and analysis.
  6. [6]
    [PDF] Introduction to Fourier transform spectroscopy - GovInfo
    Fourier transform spectroscopy results from the application of Fourier ... Fourier transforms maybe defined by the following two integrals: cO f ( x ).
  7. [7]
    The History and Current Status of Fourier Transform Spectroscopy
    This paper is concerned with the development of Fourier transform spectroscopy from its beginnings in Michelson's visibility technique through the present ...
  8. [8]
    [PDF] Fourier-Spectroscopy, 1950-1970 - PhilArchive
    Devised in large part by Albert Michelson in the 1890s, Fourier spectroscopy was developed after the second world war by researchers in Britain, France and ...
  9. [9]
    [PDF] CORNELL UNIVERSITY - NASA Technical Reports Server (NTRS)
    a spectrometer goes back to the original work of Golay (1949) and Fellgett ... In the Fourier transform each spectral element can be viewed as having a ...
  10. [10]
    In Search of Space: Fourier Spectroscopy 1950–1970 | SpringerLink
    Early history of Fourier transform spectroscopy. Infrared Physics 24: 69–93 ... Fellgett, Peter (1951). Theory of infra-red sensitivities and its ...
  11. [11]
    Colorful Solar Instrument Treks East for a New Career - NSO
    The FTS was built in 1971-76 and used almost continuously through 2001 at NSO's McMath-Pierce Solar Telescope at Kitt Peak for solar and atmospheric physics, ...
  12. [12]
  13. [13]
    [PDF] nuclear magnetic resonance fourier - transform - Nobel Prize
    This Nobel lecture provides a glimpse behind the scene of an NMR toolmaker's workshop. Nuclear spin systems possess unique properties that predestinate them for ...
  14. [14]
    Recent applications and current trends in analytical chemistry using ...
    Synchrotron radiation based Fourier-transform infrared (SR-FTIR) microspectroscopy is an emerging technique, which is increasingly employed in analytical ...
  15. [15]
    [PDF] 8. Fourier transform spectroscopy - Michelson interferometer (revisited)
    The Michelson interferometer measures the spectrum. The Fourier transform calculates the spectrum from the measured interferogram, relating time domain signals ...
  16. [16]
    [PDF] Introduction to Fourier transform spectroscopy - GovInfo
    In symbolic notation J- stands for "the Fourier transform of ...". Thus, for example, ^IF(x)} = F(0 means the Fourier transform of f(x) is ...
  17. [17]
    [PDF] Fourier Transform Infrared Spectroscopy for the Measurement of ...
    May 23, 2012 · This equation shows that the observed spectrum B(σ) is real and even, if the interferogram is an even function. Only the side of the ...
  18. [18]
    Fourier Transform Infrared Spectrometry - jstor
    Oct 21, 1983 · The Beckman FT-1000 and 2000 series FTIR spectrometers are one ex- ample of several inexpensive Fourier spectrometers that are now commercial-.
  19. [19]
    Apodizing functions for Fourier transform spectroscopy
    Apodizing functions are used in Fourier transform spectroscopy (FTS) to reduce the magnitude of the sidelobes in the instrumental line shape (ILS), ...
  20. [20]
    Fourier Transform Spectroscopy - RP Photonics
    Fourier transform spectroscopy is a method where one computes an optical spectrum from raw data by applying a Fourier transform algorithm.What is Fourier Transform... · Operation Principle · Calibration of Spectral Power...
  21. [21]
    FT-IR Spectroscopy - Newport
    The process of conversion is through the Fast Fourier Transform algorithm. The discovery of this method by J.W. Cooley and J.W. Tukey in 1965, followed by an ...
  22. [22]
    [PDF] Fourier Transform Interferometry - DTIC
    Oct 29, 1999 · The basic Michelson interferometer, as used for Fourier transform spectroscopy, is analyzed in this report. The principles of operation are ...
  23. [23]
    What's the Appropriate Infrared Detector for FTIR Analysis and How ...
    Dec 17, 2024 · Thermoelectrically cooled MCT detectors offer significant advantages over traditional IR detectors, combining high speed with operational convenience.
  24. [24]
    None
    ### Summary of Spectrum Acquisition and Processing in FTIR
  25. [25]
    Data acquisition and interferogram data treatment in FT-IR ...
    Traditional FT-IR spectrometers use the zero-crossings of an interferogram from a control HeNe laser as a trigger signal for digitizing the analog signal of the ...
  26. [26]
    The FFT in FTIR - Essential FTIR
    The Fast Fourier Transform (FFT) applied to FTIR Data · Normalize the Interferogram · Zero-Fill the Interferogram · Prepare Phase-Correction Data · The sample data ...Missing: acquisition | Show results with:acquisition
  27. [27]
    Removing aperture-induced artifacts from fourier transform infrared ...
    Two Fourier transform infrared intensity artifacts have been observed at moderately high (0.1 cm-1) spectral resolution: Light reflected off the aperture ...
  28. [28]
    Spectral Subtraction | Spectroscopy Online
    May 1, 2021 · Spectral subtraction allows users to remove the spectral contribution of known components from a mixture spectrum so the spectra of unknown components can be ...Missing: ghosting aliasing atmospheric
  29. [29]
    Fourier Transform and Apodization - Shimadzu
    Applying some type of function to Fourier transform integration to reduce the ripples, as in this example, is called "apodization" and the function is known as ...
  30. [30]
    [PDF] Advanced FTIR Spectroscopy - Nicolet CZ
    with an MCT detector at a data collection speed of 77 spectra/sec.1. Chapter 2: High-Resolution, High-Speed and. Spectral Range Extension of FTIR. 2.1 High ...
  31. [31]
  32. [32]
    THz Time-Domain Spectroscopy
    In order to create THz pulses, a diode-pumped CW laser at 532 nm pumps a mode-locked Ti:sapphire crystal to create ultra short (100 fs or shorter) optical ...
  33. [33]
    NMR Basics
    Shimming is adjusting the homogeneity of the magnetic field, so that every part of the sample in the NMR tube experiences exactly the same field strength. OK, ...Missing: 1-100
  34. [34]
    Mid-infrared supercontinuum-based Fourier transform spectroscopy ...
    Jun 10, 2022 · Broadband time-resolved absorption and dispersion spectroscopy of methane and ethane in a plasma using a mid-infrared dual-comb spectrometer.
  35. [35]
    Guidelines for the Use of Oversampling in Protein NMR
    One is that the oversampling increases the dynamic range, the ratio between the largest and smallest signals that can be detected [3,4].
  36. [36]
    [PDF] A versatile ultrastable platform for optical multidimensional Fourier ...
    In the single-quantum 2DFT spectra, cycling the phase of the excitation pulses during a. 2DFT scan suppresses the noise along the diagonal. In NMR spectroscopy, ...
  37. [37]
    Analysis of Wedge Error of Beam Splitter in Spatial Modulation ...
    Due to the restriction of manufacture precision to beam splitter in spatial modulated Fourier transform infrared spectrometer, the wedge error will be ...
  38. [38]
  39. [39]
    Fourier Transform Spectroscopic Imaging Using an Infrared Focal ...
    A wide-field imaging approach for simultaneous super-resolution surface-enhanced Raman scattering bioimaging and spectroscopy.
  40. [40]
    Channel dispersed Fourier transform spectrometer - Nature
    Jul 12, 2018 · A stationary channel dispersed Fourier transform (CDFT) spectrometer that efficiently utilizes two-dimensional imagers to obtain high-resolution input spectra.
  41. [41]
    Stationary Fourier-transform spectrometer - Optica Publishing Group
    The stationary interferometer is fabricated without a beam splitter or moving components. The measured interferogram visibilities, which are limited by the size ...
  42. [42]
    Portable Fourier transform infrared spectroradiometer for field ...
    A hand-held, battery-powered Fourier transform infrared spectroradiometer weighing 12.5 kg has been developed for the field measurement of spectral radiance ...
  43. [43]
    A Sagnac Fourier spectrometer
    ### Summary of Sagnac Fourier Spectrometer from https://opg.optica.org/abstract.cfm?uri=oe-25-8-a447
  44. [44]
  45. [45]
  46. [46]
  47. [47]
  48. [48]
    Photon-noise-limited Fourier spectroscopy
    ### Summary of Photon-Noise-Limited Case for Fourier Spectroscopy
  49. [49]
    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.
  50. [50]
    Spectral noise due to sampling errors in Fourier-transform ...
    Abstract. An assessment is made of the spectral noise in Fourier-transform spectroscopy caused by sampling errors in the interferogram acquisition.Missing: limitations instabilities load
  51. [51]
    Research on Fourier Transform Spectral Phase Correction Algorithm ...
    This paper proposes an algorithm called the Cross-Teager–Kaiser ψ B Energy Operator–Nonlinear Calibration Model (CTKB-NCM) for phase error correction.
  52. [52]
    Evaluation of the sensitivity to mechanical vibrations of an IR Fourier ...
    Dec 27, 2005 · This article presents an experimental and theoretical investigation of the sensitivity to mechanical disturbances of a Fourier-transform ...
  53. [53]
    Application of open-path Fourier transform infrared spectroscopy ...
    Open-path Fourier transform infrared spectroscopy (OP-FTIR) has often been used to measure hazardous or trace gases from hot point sources (e.g. volcano, ...
  54. [54]
    Enhancing spatial resolution in Fourier transform infrared spectral ...
    Dec 20, 2023 · By combining unsupervised machine learning algorithms such as SVD and NMF, the effective noise reduction can be achieved through dimensionality ...
  55. [55]
    12.8: Infrared Spectra of Some Common Functional Groups
    Jan 1, 2025 · Use an infrared spectrum to determine the presence of functional groups, such as alcohols, amines and carbonyl groups, in an unknown compound.Missing: gas liquid cells
  56. [56]
    [PDF] Organic Compounds FT-IR Spectroscopy - IntechOpen
    Jun 29, 2012 · The IR absorption spectra can be recorded for solid, liquid or gaseous samples. The most common presentation state of samples in drug ...
  57. [57]
    [PDF] Impact of active gas purge on FTIR measurements
    The results unequivocally demonstrate that gas purge is a simple yet highly effective and efficient method to remove water vapor and CO₂ from FTIR ...Missing: polyethylene | Show results with:polyethylene
  58. [58]
    Polyethylene Sheets for FAR IR SAMPLING
    Polyethylene is a useful transmission material for far infrared sampling. Polyethylene is non-absorbing between 16 and 2500 microns. Transmission is about 80%.
  59. [59]
    Superconducting bolometer for far-infrared Fourier transform ...
    Jan 20, 2003 · A voltage-biased superconducting bolometer with a heat sink temperature of 4.2 K has been developed for Fourier transform spectroscopy in ...Missing: rotational | Show results with:rotational
  60. [60]
    [PDF] Detection of the 267 GHz J = 1-0 Rotational Transition of PH3 in ...
    Dec 14, 1993 · Far-infrared observations can detect energetic rotational transitions. (Haas et al. 1985, 1986), but measurements at these wavelengths have thus ...
  61. [61]
    [PDF] Method TO-16 - Long-Path Open-Path Fourier Transform Infrared ...
    5.14 Monitoring path—the actual path in space over which the pollutant concentration is measured and averaged. 5.15 Monostatic System—a system with the source ...
  62. [62]
    [PDF] Method 320 - U.S. Environmental Protection Agency
    Aug 2, 2017 · 1.1. 1 Analytes. Analytes include hazardous air pollutants (HAPs) for which EPA reference spectra have been developed. Other compounds can also ...
  63. [63]
    IR-EcoSpectra: Exploring sustainable ex situ and in situ FTIR ... - NIH
    FTIR became a typical approach for identifying unknown substances in forensic investigations and drug analyses in the 1980s [10]. The invention of micro-FTIR ...
  64. [64]
  65. [65]
    Direct Detection and Characterization of Exoplanets Using Imaging ...
    Space-based direct imaging provides prospects for detection and spectral characterization of exoplanets at optical and near-infrared wavelengths.
  66. [66]
    [PDF] Development and Results of a New Generation Imaging Fourier ...
    Feb 16, 2024 · With its low to high dynamic resolution capability, the IFTS is positioned to be a powerful tool to perform cutting-edge exoplanet science. 2.
  67. [67]
    [PDF] Hig h-Resolution Broadband Spectral Interferometry - OSTI
    The ED1 is a hybrid instrument sharing advantages of both purely dispersive, and purely interferometric systems such as the Fourier Transform Spectrometer (FTS) ...<|separator|>
  68. [68]
    SITELLE: an Imaging Fourier Transform Spectrometer for the ...
    An astronomical iFTS is basically a Michelson interferometer inserted into the collimated beam of an astronomical camera system equipped with two detectors.Abstract · INTRODUCTION · THE INSTRUMENT · INSTRUMENT CAPABILITIES...
  69. [69]
    TCCON - Total Carbon Column Observing Network
    TCCON is a network of ground-based Fourier Transform Spectrometers recording direct solar spectra in the near-infrared spectral region.Missing: remote sensing
  70. [70]
    The Total Carbon Column Observing Network - Journals
    May 28, 2011 · The Total Carbon Column Observing Network (TCCON) is a ground-based network of Fourier transform spectrometers (FTSs) designed to retrieve ...
  71. [71]
    Ultra-broadband coherent open-path spectroscopy for multi-gas ...
    Mar 17, 2025 · Differential optical absorption spectroscopy (DOAS) is a classical technique in open-path spectroscopy [[16], [17], [18]] using sunlight or a ...
  72. [72]
    Monitoring Urban Greenhouse Gases Using Open-Path Fourier ...
    Here, we present an Open-Path Fourier Transform Infrared (OP-FTIR) spectroscopy system for GHG monitoring in Toronto, Canada.Missing: DOAS | Show results with:DOAS
  73. [73]
    An open-path observatory for greenhouse gases based on ... - AMT
    Dec 20, 2023 · The observatory measures path-averaged concentrations of CO 2 and CH 4 along a 1.55 km path in the urban boundary layer above the city.
  74. [74]
    Fourier transform spectrometer for Greenhouse Gases Observing ...
    The Greenhouse gases Observing SATellite (GOSAT) is designed to monitor the global distribution of carbon dioxide (CO2) from the space. GOSAT is a joint project ...Missing: spectroscopy | Show results with:spectroscopy
  75. [75]
    Fourier transform spectrometer on GOSAT and GOSAT-2
    Nov 17, 2017 · The measurement targets of the GOSAT is to observe CO2 and CH4 column density with relative accuracy of 1 % for CO2 and 2 % for CH4 at 1,000 km ...Missing: spectroscopy | Show results with:spectroscopy
  76. [76]
    Spaceborne hyperspectral imaging with a static Fourier transform ...
    Jan 24, 2013 · Imaging static Fourier transform spectrometers could sharpen hyperspectral data acquisition on small satellites.Missing: spectroscopy | Show results with:spectroscopy
  77. [77]
    Impact of image registration errors on the quality of hyperspectral ...
    Feb 13, 2024 · Imaging static Fourier transform spectrometry (isFTS) is used for pushbroom airborne or spaceborne hyperspectral remote sensing. In isFTS, a ...
  78. [78]
    Update on the GOSAT TANSO–FTS SWIR Level 2 retrieval algorithm
    Mar 22, 2023 · The retrieval results from the updated algorithm showed improvements in fitting accuracies in the O 2 A, weak CO 2 , and CH 4 bands of TANSO–FTS.