Fact-checked by Grok 2 weeks ago

Monochromator

The monochromator's origins trace back to the mid-19th century with the spectroscope invented by and in 1859 for . Modern grating-based designs, such as the Czerny–Turner configuration, were developed in 1930 by Max Czerny and Arthur Francis Turner. A monochromator is an optical device that selects a narrow range of wavelengths from a polychromatic source, effectively producing monochromatic for applications in and analysis. It functions as a tunable , allowing precise control over the center and of the transmitted . The working principle of a monochromator relies on the of light into its constituent wavelengths using either prisms or , followed by spatial separation and selection via slits. In prism-based systems, light passes through a refractive medium where shorter wavelengths bend more than longer ones due to varying refractive indices, though this results in non-linear . monochromators, more common in modern instruments, use periodic grooves to diffract light via constructive , with the grating equation m\lambda = d(\sin \theta_i + \sin \theta_m) determining the angles for different orders m and wavelengths \lambda, where d is the groove spacing. Key components include an entrance slit to define the input beam, collimating and focusing mirrors, the dispersive element, and an exit slit to isolate the desired wavelength band, enabling resolutions as fine as 0.01–0.1 nm. Common types of monochromators include the Czerny-Turner design, which employs two concave mirrors for high image quality and is widely used in UV-visible spectroscopy, and the Fastie-Ebert configuration, featuring a single spherical mirror for cost-effective but aberration-prone operation. Holographic gratings with 1200–1400 grooves per mm are standard for UV-visible ranges, offering high up to 60,000 while minimizing scattered light compared to older ruled gratings. These instruments are essential in fields like and , where they enable precise wavelength scanning and serve as alternatives to tunable lasers for generating narrow-bandwidth illumination.

Overview

Definition and Purpose

A monochromator is an that produces nearly monochromatic light by dispersing a broadband light source into its spectral components and selecting a narrow band of wavelengths centered at a desired value. This process isolates of effectively a single color or extremely narrow from polychromatic input, enabling precise control over the output . The primary purpose of a monochromator is to provide tunable, illumination for applications in optical , where pure spectral lines are crucial for accurate analysis. It facilitates the identification of chemical compositions by exciting samples at specific wavelengths, enhances signal-to-noise ratios in detection systems by minimizing , and supports wavelength-dependent experiments such as or measurements. In these contexts, monochromators serve as essential components in spectrophotometers and other analytical tools, allowing researchers to probe material properties with high specificity. Unlike optical filters, which offer fixed selection with wider passbands (typically 10–250 ), monochromators provide adjustable for greater precision and versatility in tuning. Whereas spectrometers disperse and display an entire for comprehensive , monochromators emphasize the of a single narrow band without full spectral presentation. Monochromators first emerged in 19th-century as part of early efforts to resolve into discrete wavelengths.

Historical Development

The development of monochromators began in the mid-19th century with prism-based designs for spectroscopic analysis. In 1856, Scottish physicist William Swan introduced an astronomical spectroscope incorporating a , which significantly improved the isolation of narrow spectral bands from broadband light sources, marking an early milestone in monochromator technology. Earlier, Joseph von Fraunhofer's 1814 work on mapping over 500 dark lines in the solar spectrum using prisms laid foundational techniques for spectral dispersion and analysis. and Robert Bunsen's contributions in 1859–1860 further advanced prism spectroscopes for identification through emission and lines, establishing as a quantitative tool. The late 19th and early 20th centuries saw the transition to , enhancing resolution and efficiency over prisms. In 1882, Henry Augustus Rowland invented the concave , enabling high-resolution solar spectrum mapping with reduced astigmatism and a ruling engine capable of 14,438 grooves per inch. advanced ruled grating production in the 1890s through interferometric precision, though refinements to ruling engines continued into the 1920s, improving groove uniformity for broader spectral applications. Robert W. Wood's 1910 invention of the optimized diffraction efficiency by angling grooves to direct more light into specific orders, with widespread adoption in the 1930s–1950s for and . The Czerny-Turner mounting, introduced in 1930 by M. Czerny and A. F. Turner, became a standard configuration for plane gratings in monochromators, using symmetrical off-axis spherical mirrors for compact, high-throughput designs. By the , computer-controlled variants of the Czerny-Turner emerged, enabling automated scanning and precise in instruments. Commercial adoption accelerated in the , with monochromators integrated into spectrophotometers for routine analytical use, such as the 1966 introduction of laser-excited Raman systems featuring dispersive gratings. Post-1980 developments included holographic gratings, first commercialized in 1969 but refined in the 1980s for lower and higher groove density via interferometric recording, reducing manufacturing defects in high-resolution monochromators. Acousto-optic tunable filters (AOTFs) gained prominence as solid-state alternatives post-1980, using ultrasonic waves in birefringent crystals for rapid, non-mechanical selection, as reviewed in 1981 applications for chemical . After 2000, monochromators increasingly incorporated fiber optics for compact, fiber-to-fiber coupling and integration with tunable lasers, enhancing portability in and sensing systems.

Operating Principles

Basic Components and Light Path

A monochromator consists of several core optical components that work together to isolate a specific wavelength from broadband light. The entrance slit serves as the initial aperture, defining the spatial extent of the incoming beam and limiting the amount of light entering the system to improve spectral purity. This slit is typically adjustable, with its width directly influencing the resolution by controlling the angular spread of the input light. Following the entrance slit, a collimating optic—usually a lens or mirror—parallelizes the diverging rays from the slit, creating a uniform beam that minimizes optical aberrations such as astigmatism and ensures efficient interaction with the dispersive element. The dispersive element, which can be a prism or diffraction grating, angularly separates the collimated light based on wavelength, spreading the spectrum across a focal plane. A focusing optic, often another mirror or lens, then reimages this dispersed spectrum onto the plane of the exit slit, concentrating the light rays for selection. The exit slit acts as a final filter, allowing only a narrow band of wavelengths to pass through while blocking others, thereby outputting nearly monochromatic light. These slits play a crucial role in resolution: narrower widths enhance the ability to distinguish closely spaced wavelengths by reducing the spectral bandwidth, though at the cost of lower light throughput. Collimation is essential to avoid aberrations that could blur the spectrum and degrade performance. The light path in a monochromator follows a sequential process to achieve wavelength selection. Broadband light first passes through the entrance slit, where it is spatially filtered to form a narrow beam. This beam is then collimated by the entrance optic, rendering the rays parallel and directing them toward the dispersive element. Upon incidence, the dispersive element separates the wavelengths into an angular spectrum, with shorter wavelengths deviated more than longer ones. The focusing optic collects these dispersed rays and images the spectrum linearly onto the exit slit plane, such that each position corresponds to a specific wavelength. Finally, the exit slit transmits only the desired narrow band, while the rest of the spectrum is rejected. In a simple ray diagram, rays from the source converge at the entrance slit, diverge slightly but are collimated to strike the dispersive element flatly, fan out post-dispersion, and reconverge at the exit slit for the selected wavelength. The throughput of the monochromator, or the intensity of the output light, is fundamentally limited by the slit dimensions and is proportional to the product of the slit width and height, assuming no vignetting or losses elsewhere in the system: I \propto w \times h where I is the output intensity, w is the slit width, and h is the slit height. This relationship highlights the trade-off between resolution (favored by small w) and signal strength (increased by larger w and h).

Dispersion Fundamentals

Dispersion in monochromators refers to the angular separation of different wavelengths of , enabling the isolation of specific components from polychromatic sources. This separation arises from the wavelength-dependent of with dispersive elements, where shorter wavelengths (e.g., ) and longer wavelengths (e.g., red light) deviate by different angles due to variations in or . In prisms, occurs through , governed by the material's n(\lambda), which decreases with increasing (dn/d\lambda < 0). As light passes through the prism, Snell's law dictates that the angle of varies with n, leading to greater deviation for shorter wavelengths; the angular is thus proportional to dn/d\lambda, the rate of change of with . For a prism at minimum deviation, the deviation angle \delta_{\min} is given by \delta_{\min} = 2 \sin^{-1} [n \sin(A / 2)] - A, where A is the apex angle (approximating to (n - 1)A for small angles), and d\delta/d\lambda \approx \frac{2 \sin(A / 2) \cos(A / 2)}{ \sqrt{1 - [n \sin(A / 2)]^2} } \cdot dn/d\lambda. In diffraction gratings, dispersion results from interference of light waves scattered by periodic grooves, producing maxima at angles where the path difference between adjacent grooves is an integer multiple of the wavelength. The grating equation, m\lambda = d (\sin \theta_i + \sin \theta_d), describes this, where m is the diffraction order (integer), \lambda is the wavelength, d is the groove spacing, \theta_i is the angle of incidence, and \theta_d is the angle of diffraction. To derive it, consider a plane wave incident at angle \theta_i on a grating with sinusoidal transmittance t(x) = a + b \cos(2\pi x / d). The incident field is E_{\text{in}}(x, t) = \sin(2\pi \nu t - k x \sin \theta_i), where k = 2\pi / \lambda and \nu = c / \lambda. The output field E_{\text{out}} = E_{\text{in}} \cdot t(x) expands via trigonometric identities into multiple plane waves: the zero-order at \theta_d = \theta_i, and higher orders where \sin \theta_{d,m} = \sin \theta_i + m \lambda / d, rearranging to the grating equation. This shows that for fixed m, d, and \theta_i, \theta_d increases with \lambda, spatially separating wavelengths. Factors influencing dispersion include material properties in prisms, such as the dispersion curve of the glass (e.g., higher |dn/d\lambda| in flint glass yields greater separation), and groove density in gratings, where smaller d (higher lines per mm) increases angular dispersion d\theta_d / d\lambda = m / [d \cos \theta_d]. In spectroscopy, wavelengths are often related to photon energy via E = hc / \lambda, where h is Planck's constant ($6.626 \times 10^{-34} J s) and c is the speed of light ($3.00 \times 10^8 m/s); for visible light, energies range from about 1.65 eV (red, 750 nm) to 3.10 eV (violet, 400 nm), facilitating comparisons in electronvolts (eV) or nanometers (nm).

Dispersive Elements

Prisms

Prism-based monochromators utilize the refractive dispersion of light through a transparent medium to separate wavelengths. The principle relies on the wavelength-dependent refractive index n(\lambda) of the prism material, which causes light rays of different wavelengths to deviate at varying angles upon refraction. This angular deviation \theta for a prism with apex angle A is given by \theta = (n-1)A for small angles in the minimum deviation configuration. The dispersion D, defined as the rate of change of deviation angle with wavelength, follows D = \frac{d\theta}{d\lambda} \propto \frac{dn}{d\lambda}, enabling spatial separation of spectral components along the focal plane. Common materials for prisms in monochromators are selected based on their dispersion properties and transmission windows to minimize absorption. Flint glass, with its high dispersion due to elevated \frac{dn}{d\lambda}, is widely used for the visible range (approximately 400-700 nm), offering effective separation in that spectrum. Quartz (fused silica) extends utility into the ultraviolet region down to about 200 nm, where it provides strong dispersion (e.g., >1 mrad nm⁻¹ near 250 nm for a 60° prism). For infrared applications up to around 2000 nm, fluorite (calcium fluoride) is employed, though all materials impose absorption limits that restrict the overall operational range to typically 200-2000 nm. Design configurations for prism monochromators emphasize simplicity and efficiency in path. The equilateral , with all angles at 60°, is a standard choice for balanced deviation and in visible and UV setups. The Littrow configuration integrates the with a reflecting surface, allowing the dispersed to retrace its path for compact single-element , commonly used in instruments operating above 200 nm. These designs often require fixed orientations for operation, with selection achieved via rotating the or exit slit. Prisms offer several advantages as dispersive elements, particularly in their straightforward construction without periodic structures or moving parts in fixed setups, which simplifies manufacturing and maintenance. They provide high throughput of radiation in the visible range due to efficient and lack of orders, enhancing in low-light applications such as . However, prism monochromators have notable limitations that constrain their use. The nonlinear dispersion, where resolution improves at shorter wavelengths but varies unevenly across the spectrum, necessitates calibration aids like cams for accurate wavelength readout. Thermal instability arises from temperature-dependent changes in n, potentially shifting the spectrum and requiring temperature control. Additionally, the limited wavelength range due to material absorption and the overall heavier, more costly construction compared to alternatives restrict their application to specific spectral regions.

Diffraction Gratings

Diffraction gratings function as dispersive elements in monochromators through arising from a periodic array of grooves on a reflective or transmissive , which diffracts incoming polychromatic into discrete spectral orders separated by . The groove periodicity determines the angular dispersion, with peak diffraction efficiency occurring at a specific blaze in optimized designs, enabling selective concentration of in desired orders for spectroscopic applications. Gratings are primarily categorized into ruled and holographic types, each produced via distinct methods that influence their performance characteristics. Ruled gratings feature grooves mechanically scribed into the surface using a , resulting in profiles such as triangular or trapezoidal shapes that can be precisely controlled for blazing. In contrast, holographic gratings are fabricated by recording patterns from beams in a photosensitive , yielding smooth sinusoidal groove profiles with inherently lower imperfections. Groove densities for these gratings typically range from 600 to 2400 lines per millimeter, tailored to cover ultraviolet-visible-infrared spectral regions from approximately 10 nm to 100 μm. Blazed gratings, a subtype commonly used in ruled designs, incorporate a sawtooth-like groove profile to redirect diffracted light constructively into a single predominant order, enhancing overall by mimicking from the groove facets. The \beta, defined as the tilt of the primary facet relative to the grating , is selected such that \sin \beta = \frac{\lambda_{\text{[blaze](/page/Blaze)}}}{2d}—where \lambda_{\text{blaze}} is the and d is the groove spacing—to achieve maximum in the under Littrow mounting conditions. This design concentrates up to 80-90% of incident energy at the , though drops outside a of roughly 20-50% around it. Key advantages of diffraction gratings include their linear wavelength dispersion, which simplifies and scanning in monochromators compared to nonlinear refractive elements, alongside broad operational spectral coverage from deep to mid-infrared. Additionally, the selectable nature of orders allows flexibility in balancing resolution and throughput, with higher orders providing finer separation at the cost of reduced intensity. Despite these benefits, diffraction gratings are susceptible to limitations such as ghosting and stray light, which arise from periodic ruling errors in ruled types or residual scattering in holographic ones, potentially degrading spectral purity. The resolving power, defined as R = \frac{\lambda}{\Delta \lambda} = m N—where m is the diffraction order and N is the total number of illuminated grooves—quantifies the minimum resolvable wavelength difference, but imperfections can reduce effective N and introduce artifacts. Volume holographic gratings represent an advanced variant, where the periodic structure is embedded as refractive index modulations within a bulk photosensitive medium, such as dichromated , offering diffraction efficiencies exceeding 70% across narrow bands while minimizing surface and ghosts. These gratings excel in high-resolution applications, achieving R > 50,000 with groove densities over 6000 lines/mm in double-pass configurations, and their protected between glass substrates enhances durability against environmental factors. Echelle gratings, a specialized blazed design with low groove densities (31-316 lines/mm) and steep blaze angles exceeding 45°, operate in high orders (e.g., m > 10) to deliver exceptional resolving powers up to 250,000, ideal for broadband, high-fidelity spectroscopy in astronomical and analytical instruments. By employing cross-dispersion with a secondary element, echelles map multiple overlapping orders into a two-dimensional format, enabling simultaneous coverage of wide spectral ranges (e.g., 350-5300 nm) with uniform efficiency above 40%.

Monochromator Configurations

Single Monochromator Designs

Single monochromator designs feature a single dispersive element, typically a , combined with focusing to isolate a narrow band from input light. These configurations prioritize simplicity and efficiency for standard spectroscopic tasks, with light entering through an entrance slit, undergoing , and exiting via a selectable exit slit or detector plane. The primary layouts—Czerny-Turner, Fastie-Ebert, and Littrow—differ in their use of mirrors and grating placement, influencing compactness, alignment ease, and aberration control. The Czerny-Turner configuration employs two separate concave mirrors flanking a plane . Incoming light from the entrance slit is collimated by the first mirror onto the grating, which disperses it by ; the second mirror then focuses the selected onto the exit slit. This allows independent adjustment of mirror positions to optimize the spectral field, providing good correction for aberration at a central wavelength while maintaining a relatively flat focal plane. However, it introduces unavoidable and across the spectrum, limiting performance at off-design wavelengths. In contrast, the Fastie-Ebert configuration uses a single large spherical mirror for both collimation and focusing, with the plane positioned midway. Light from the curved entrance slit reflects off the mirror to the , diffracts, and returns to the same mirror for refocusing onto a curved exit slit in the same plane. This shared-mirror approach simplifies construction and alignment by reducing component count, making it more compact than the Czerny-Turner. Drawbacks include increased from direct reflections and reduced flexibility in correcting aberrations like and , as the fixed mirror-to-slit geometry constrains design parameters. The Littrow configuration achieves autocollimation by aligning the incident and diffracted beams along the same path, typically using the itself or a single auxiliary mirror for focusing. enters near the , diffracts back toward the entrance slit (with a slight offset for separation), and is collected accordingly. This design maximizes efficiency at the and minimizes optical elements for utmost compactness. Limitations arise from out-of-plane aberrations when slits are offset and mechanical constraints on rotation, restricting the accessible range without efficiency losses. Geometrical design in these single-stage systems involves key trade-offs between and slit separation to balance , throughput, and size. Increasing the enhances spectral by enlarging the dispersed image scale and reducing angular aberrations, but it enlarges the overall instrument and raises costs due to larger . Conversely, greater separation between entrance and exit slits improves by allowing finer wavelength selection, yet it amplifies off-axis aberrations such as in non-symmetric layouts like Czerny-Turner. In the paraxial approximation, the position of the spectral image at the exit focal plane is determined by the focusing , approximated as x \approx f \beta, where f is the of the camera mirror and \beta is the diffraction angle from the (with higher-order terms neglected for small angles). This relation guides slit placement to capture the desired while minimizing defocus. These designs offer advantages in compactness and cost-effectiveness, enabling straightforward implementation for routine where ultra-high purity is not required. They typically operate at f-numbers from f/4 to f/8, providing a practical compromise between light-gathering power and aberration control. A notable limitation is elevated levels from ghosts, mirror scatter, and slit edge effects, which can degrade signal-to-noise in low-light conditions. Most commercial single monochromators adopt the Czerny-Turner layout, delivering resolutions of 0.1–10 nm depending on groove density, , and slit width.

Double and Multiple Monochromators

Double monochromators incorporate two dispersive stages in series, with the exit slit of the first serving as the entrance slit for the second, to achieve superior spectral selection compared to single monochromators. This configuration significantly enhances rejection and , making them suitable for applications requiring high spectral purity. Multiple monochromators extend this principle with three or more stages for even greater performance in extreme conditions. Additive double monochromators combine the dispersions of both stages, effectively doubling the reciprocal linear dispersion and thereby improving resolution; for example, two identical grating-based stages can reduce the spectral bandwidth by a factor of approximately 2 relative to a single stage of equivalent design. In contrast, subtractive double monochromators configure the second stage to counter the dispersion of the first, resulting in near-zero net dispersion at the exit and a spectrally uniform output beam that excels at rejecting off-wavelength light; examples include pairing a in the first stage with a in the second. These systems typically employ a tandem serial design, where stages are aligned sequentially, though some advanced models allow switching between additive and subtractive modes via adjustable optics. Subtractive configurations are particularly prevalent in setups for effective notch filtering of excitation wavelengths. In fluorescence microscopy, double monochromators enable precise isolation of narrow emission lines by minimizing interference from broadband sources. The primary advantages include bandwidth narrowing by factors of 2 or more and stray light suppression to levels below $10^{-10}, achieved through the multiplicative rejection of each stage. However, these benefits come with drawbacks such as reduced optical throughput due to additional absorptive and losses, greater mechanical complexity in alignment and operation, and higher manufacturing costs compared to single-stage designs. Multiple monochromators amplify these traits, offering stray light rejection potentially exceeding $10^{-12} in triple configurations, but at proportionally increased throughput penalties and system intricacy.

Performance Characteristics

Spectral Resolution and Bandwidth

Spectral resolution in a monochromator refers to its ability to distinguish between two closely spaced wavelengths, quantified by the resolving power R = \lambda / \Delta\lambda, where \lambda is the central wavelength and \Delta\lambda is the smallest resolvable wavelength difference. This metric is fundamental to the instrument's performance, as higher R values enable finer spectral detail in applications like emission line analysis. The overall resolution is limited by both intrinsic and extrinsic factors. Intrinsic resolution is determined by the dispersive element itself, independent of mechanical components like slits. For diffraction gratings, the theoretical maximum resolving power follows the Rayleigh criterion, which states that two wavelengths are just resolvable when the maximum of one pattern coincides with the first minimum of the other, yielding R = mN, where m is the and N is the total number of illuminated grooves on the . This establishes the grating's inherent limit, often approaching R \approx 10^5 to $10^6 for high-quality ruled gratings with thousands of grooves. Extrinsic resolution, in contrast, is imposed by the instrument's geometry, particularly the slits, and typically dominates in practical setups. The effective spectral bandwidth \Delta\lambda at the exit slit represents the wavelength range passed through the monochromator for a given setting, which directly limits the achievable resolution when slit-limited. To derive this, consider the light path: polychromatic input is collimated, dispersed by the grating into angularly separated wavelengths according to the grating equation m\lambda = d (\sin\alpha + \sin\beta), where d is the groove spacing, \alpha the incidence angle, and \beta the diffraction angle. The angular dispersion is D = \frac{d\beta}{d\lambda} = \frac{m}{d \cos\beta}. At the focal plane of length f, the linear position x = f \tan\beta \approx f \beta (for small angles), so the linear dispersion is \frac{dx}{d\lambda} = f D, or reciprocally, \frac{d\lambda}{dx} = \frac{1}{f D}. For an exit slit width w (or the magnified image of the entrance slit, whichever is larger), the corresponding wavelength spread is the product of the physical width and the dispersion rate: \Delta\lambda = w \cdot \frac{d\lambda}{dx} = \frac{w}{f D}. Thus, the slit-limited resolving power is R = \frac{\lambda}{\Delta\lambda} = \frac{\lambda f D}{w}, highlighting the trade-off between resolution (narrower w, longer f, higher D) and light throughput. In traditional scanning monochromators with point detectors like photomultiplier tubes, the exit slit width sets the , often adjustable from 0.1 nm to tens of nm depending on the (e.g., 0.3 m instruments achieve ~0.05 nm at narrow slits). However, modern imaging spectrographs, which pair fixed with () detectors, extend resolution limits beyond analog slits by eliminating the exit slit entirely. Here, the effective bandwidth is determined by the of the slit image width, grating envelope, and pixel sampling; pixel sizes of 10–25 μm can yield resolutions down to the grating's intrinsic mN limit (e.g., \Delta\lambda \approx 0.01 nm at 500 nm for high-groove-density gratings), provided the dispersion maps multiple pixels per resolution element per the . This pixel-limited regime enables higher without sacrificing throughput, as seen in array-based systems resolving in spectra.

Stray Light and Dynamic Range

Stray light in monochromators refers to unintended light that reaches the detector, compromising purity and measurement accuracy. Primary sources include scattered light from dust, surface imperfections, or optical roughness on mirrors and ; off-order , where higher or lower orders overlap the desired ; and ghosts, which are faint spurious lines arising from periodic errors in ruled grating groove spacing. These contributions can broaden peaks or introduce baseline elevation, particularly in high-resolution systems like Czerny-Turner configurations. Rejection of stray light is achieved through optical design elements such as baffles, which absorb or redirect scattered rays to minimize while blocking off-axis paths, and double monochromator stages, where a second dispersive element further filters unwanted light, often reducing levels multiplicatively. Typical stray light levels in single monochromators are below 10^{-4} relative to the signal, with high-quality holographic gratings achieving around 1.5 \times 10^{-5} in visible regions, though values can reach 0.1% without mitigation. This stray light directly impacts the , defined as DR = 10 \log (I_{\max} / I_{\min}), where I_{\max} and I_{\min} are the maximum and minimum measurable intensities; in stray-limited cases, the effective range approximates -10 \log S, with S as the stray light fraction, limiting the instrument's ability to resolve low signals against high backgrounds alongside detector noise. Additional mitigation strategies include order-sorting filters, which block specific unwanted orders by transmitting only the desired wavelength band, and tilted slits or detectors to prevent re-entrant reflections that exacerbate ghosts and . In UV-Vis , stray light typically limits accurate measurements to below 2 AU, as higher values cause nonlinear deviations due to the stray contribution dominating the transmitted signal. Double monochromators enhance rejection for such applications, often achieving stray levels below 10^{-8}.

Wavelength Range and Dispersion

Monochromators are engineered to cover a broad spectrum of wavelengths, from (UV) through (IR), with operational limits primarily dictated by the transmission characteristics of optical materials and atmospheric absorption. In the UV and visible (Vis) regions, fused silica (quartz) enable transmission starting from approximately 180 nm, where the material's occurs due to increasing absorption below this wavelength. For air-path systems, the practical lower limit is around 185 nm owing to oxygen absorption in the atmosphere. The visible range spans 400–700 nm without significant material constraints, while extension into the near-IR reaches up to 2.5 μm using the same quartz components. In the mid-IR, specialized materials like (CaF₂), which transmits up to 8 μm, or (KBr), effective to 25 μm, allow coverage from 2.5 μm to 25 μm, aligning with fundamental vibrational needs. Beyond 25 μm, far-IR operation faces challenges from material opacity and requires alternative designs, such as windows, but upper limits can extend to 40 μm in vacuum-purged systems to mitigate absorption. These ranges ensure compatibility with diverse sources, from lamps in UV to globar sources in IR. Dispersion quantifies the spatial separation of wavelengths in a monochromator, expressed as angular dispersion d\theta / d\lambda (in radians per unit wavelength) or linear dispersion d\lambda / dx (wavelength change per unit distance at the focal plane). For diffraction grating-based systems, angular dispersion varies with the diffraction order m and is higher at larger , enhancing separation but complicating optical alignment. In prism-based monochromators, is inherently nonlinear, as the refractive index's wavelength dependence causes uneven spacing—greater in the UV and diminishing toward the —unlike the more uniform behavior of gratings at low . The dispersion for gratings derives from the grating equation m\lambda = d (\sin i + \sin \theta), where d is the groove spacing and i is the fixed incidence . Differentiating with respect to \lambda yields: \frac{d\theta}{d\lambda} = \frac{m}{d \cos \theta} This relation shows that increases with order m and groove density $1/d, but decreases with the cosine of the diffraction \theta. Linear dispersion then follows as d\lambda / dx = (d\theta / d\lambda)^{-1} \times f, where f is the , tying spread to physical . Achieving high dispersion improves wavelength isolation but introduces trade-offs: it compresses the spectrum into a narrower angular field, potentially limiting the instantaneous coverage without mechanical scanning, or requires larger apertures to maintain throughput across the full range. Scanning mechanisms, such as grating rotation, mitigate this by sequentially accessing wavelengths, though at the cost of acquisition time for broad surveys. For (EUV) and applications, where wavelengths fall below 10 , conventional prisms and ruled gratings are impractical due to and fabrication limits; instead, single-crystal monochromators, often from the phthalate family like phthalate (KAP) with a 2d spacing of 26.6 , provide via Bragg over 2.3–25 (0.23–2.5 ). These crystal systems extend access to high-energy regimes, supporting applications in diagnostics and beamlines, with governed by the lattice spacing and incidence geometry rather than grooves or .

Design Considerations

Geometrical and Optical Design

The geometrical design of monochromators involves selecting appropriate mount types to optimize performance for specific spectral requirements, such as high-order operation to achieve enhanced dispersion. Echelle gratings, characterized by coarse rulings and high blaze angles, are particularly suited for high-order mounts (), enabling compact layouts with superior resolution in applications like . In these configurations, the grating is typically mounted in a Littrow or off-plane arrangement, where the incident and diffracted beams maintain fixed angles (e.g., 2α ≈ 15°), minimizing mechanical adjustments while maximizing order separation through geometrical factors like order length (f₂ Δβ) and tilt angles influenced by cross-dispersers. Aberration correction is a critical geometrical factor, often addressed through the use of aspheric mirrors to mitigate and in fast systems. For instance, in Ebert-style monochromators with f/5 and focal lengths around 40 inches, aspherizing the collimating mirror and surfaces corrects residual aberrations after coma elimination, allowing simultaneous observation across wide spectral ranges (e.g., 1000–4000 ). This approach balances with throughput by reducing image distortion without relying on slower f-numbers. Optical design parameters like and anamorphic further influence layout optimization. The , defined as the divided by the entrance beam diameter, determines light collection efficiency; lower (e.g., f/3.4) in compact monochromators enhance throughput but introduce aberrations that must be managed. Anamorphic , arising from differing meridional and sagittal dispersions in systems, is given by the ratio of diffracted to incident beam widths (b/a), which varies with and angular configuration (α, β), necessitating designs that compensate for slit distortion. Ray tracing simulations are essential for reduction, particularly in off-axis geometries where spherical mirrors cause focal line separation; by tracing rays through the system, designers minimize via optimized mirror curvatures and tilts, ensuring stigmatic imaging. Throughput (T) in monochromator designs is fundamentally limited by étendue , expressed as T ∝ (slit area) × (optical ), where the étendue G = π S sin²Ω remains invariant across the optical train (S: area, Ω: half-angle). This conservation principle dictates that Φ = B × G, with radiance B fixed, so geometrical mismatches (e.g., oversized slits) reduce without improving . Key considerations include managing off-axis angles to prevent , where beam clipping at mirror edges reduces illumination uniformity; designs limit these angles (e.g., <15° incidence) to maintain full aperture usage. Software tools like Zemax OpticStudio enable ray tracing-based simulations to evaluate these effects, optimizing layouts for aberration balance and throughput. For high-power applications, such as synchrotron beamlines, finite element analysis (FEA) assesses thermal distortions under heat loads exceeding 100 W, guiding material selection and cooling strategies to preserve optical alignment.

Slit and Focal Length Optimization

In monochromators, the entrance slit serves to limit the spatial extent of the incoming light beam, defining the effective aperture and reducing stray light. The beam is then collimated by subsequent optics for the dispersive element. The exit slit, positioned in the focal plane after dispersion, selects the desired spectral bandwidth by allowing only a portion of the dispersed spectrum to pass through to the detector. Slit dimensions are rectangular, with the width primarily influencing spectral resolution and the height determining overall light throughput; wider slits increase throughput but degrade resolution by admitting a broader range of wavelengths, while taller slits maximize collection efficiency without impacting resolution, typically limited by the incident beam's vertical extent to avoid vignetting. The focal length f of the monochromator optics plays a critical role in performance, as longer focal lengths enhance spectral resolution since the minimum resolvable wavelength difference scales inversely with f, given by \Delta \lambda_{\min} = \frac{w}{f D}, where w is the slit width and D represents the dispersion factor (incorporating grating groove density and angular dispersion). However, extending f increases the instrument's f-number (f/\# = f / D_{\text{aperture}}), which reduces optical speed and light-gathering efficiency for a fixed aperture size, thereby lowering throughput and signal-to-noise ratio (S/N) in low-light conditions. Optimization of slits and focal length involves balancing resolution with practical throughput needs, often by setting the slit width w \approx f \times \Delta \lambda \times D to achieve a target bandwidth \Delta \lambda matched to the application's requirements, ensuring the exit slit images the entrance slit without over- or under-sampling the spectrum. This adjustment must account for the trade-off with S/N, as narrower slits for higher resolution (\Delta \lambda_{\min}) reduce photon flux, potentially necessitating longer integration times or brighter sources. In commercial instruments, variable slits—often motorized and adjustable from micrometers to millimeters—enable real-time tuning for different experiments, while slit heights are fixed or semi-adjustable to accommodate beam sizes up to several millimeters. Some designs incorporate apodized or shaped slits to produce smoother instrumental line profiles, minimizing sidelobes in the spectral response for improved accuracy in peak analysis.

Applications

In Spectroscopy Techniques

In ultraviolet-visible (UV-Vis) absorption spectroscopy, monochromators enable the scanning of wavelengths from a broadband source to measure the absorption spectra of samples, allowing the determination of concentration via the by isolating specific wavelengths and minimizing interference from polychromatic light. This isolation ensures that the absorbance, proportional to the path length and concentration of the absorbing species, accurately reflects molecular transitions without contributions from off-peak wavelengths. For fluorescence spectroscopy, monochromators are employed both for excitation wavelength selection to target specific fluorophores and for emission wavelength scanning to capture the resulting spectra, often using double monochromator configurations to reduce stray light and enhance signal-to-noise ratios. In Raman spectroscopy, double or triple monochromators are critical for rejecting Rayleigh scattered light, which is intense at the excitation wavelength, by providing high stray light rejection ratios of less than 10^{-5}, thus allowing the detection of weak inelastic scattering signals shifted by vibrational frequencies. This configuration disperses the spectrum additively across multiple stages, effectively notching out the elastic scatter while preserving Raman bands for analysis of molecular vibrations. Although Fourier-transform infrared (FTIR) spectroscopy predominantly relies on interferometry for broad spectral coverage, dispersive monochromators are used in some less common IR setups for time-resolved measurements, where sequential wavelength selection is needed to monitor transient absorptions. In such applications, the monochromator facilitates photomodulated experiments by tuning to specific IR bands, enabling the study of dynamic processes like reaction kinetics. Monochromators support continuous wavelength tuning through mechanical rotation of diffraction gratings, which adjusts the dispersion angle to select desired bands, often synchronized with optical choppers for modulation spectroscopy to distinguish sample signals from background via lock-in detection at the chopping frequency. This synchronization enhances sensitivity in low-light conditions by rejecting DC offsets. In time-resolved spectroscopy with pulsed sources, such as femtosecond lasers, specialized grating monochromators preserve pulse duration while selecting narrow bandwidths for pump-probe experiments, allowing the observation of ultrafast dynamics like electronic excitations on picosecond timescales. These designs minimize temporal broadening, ensuring high temporal resolution alongside spectral isolation for applications in transient absorption studies.

In Advanced Scientific Instruments

In advanced scientific instruments, monochromators play a critical role in high-precision applications requiring selective wavelength isolation beyond standard laboratory spectroscopy. In synchrotron radiation facilities, crystal monochromators are essential for selecting specific energies from the broad-spectrum X-ray beams produced by storage rings. These devices typically employ double-crystal configurations, such as parallel or sagittal geometries, to achieve high energy resolution while maintaining beam intensity. The operational energy range spans from approximately 0.1 keV to over 100 keV, enabling experiments in X-ray diffraction, absorption spectroscopy, and imaging at facilities like the or . In astronomical instruments, monochromators facilitate detailed spectral analysis of celestial objects by isolating narrow emission lines against broadband backgrounds. For instance, panoramic monochromators, often based on Fabry-Pérot interferometers, are integrated into solar telescopes like the THEMIS instrument on Tenerife, providing tunable narrowband imaging around lines such as H-alpha at 656.3 nm with bandwidths as low as 0.0022 nm (22 mÅ). These systems enable high-resolution studies of solar chromospheric dynamics and prominence structures by scanning across the spectrum without mechanical slits, improving throughput for faint sources. Tunable monochromators are integral to laser systems for precise output wavelength control. In dye lasers, diffraction gratings within the laser cavity serve as tuning elements, allowing continuous adjustment over broad ranges (e.g., 400-800 nm) by rotating the grating to select cavity modes, achieving linewidths below 0.01 nm. Similarly, in optical parametric oscillators (OPOs), external monochromators filter the broadband signal and idler outputs, enabling selection of specific wavelengths from tunable ranges like 410-2300 nm for applications in nonlinear optics and coherent Raman spectroscopy. In extreme ultraviolet (EUV) lithography, monochromators ensure monochromatic illumination for patterning nanoscale features. Blazed diffraction gratings in beamline monochromators, such as those at the Advanced Light Source's Beamline 12.0.1, isolate wavelengths around 13.5 nm with high efficiency (>10%) and resolutions exceeding 1000, supporting mask inspection and resist evaluation in next-generation fabrication. For advanced , monochromators enable gating in confocal systems. Dispersive elements, such as prism-grating combinations, separate spectra spatially on a detector array, allowing software-based selection of specific bands (e.g., isolating GFP at 510 nm from ) to enhance contrast and enable of biological samples. Recent advancements have focused on compact, integrated designs for portability and efficiency. Fiber-coupled monochromators, which interface directly with optical fibers, provide seamless isolation from 190 nm to 5.5 µm for and portable setups. MEMS-based tunable monochromators, utilizing electrostatically actuated blazed gratings, achieve resolutions of ~5 nm in volumes under 1 cm³, ideal for on-chip integration in handheld devices. Enhanced integration with detectors, such as arrays in spectroradiometer configurations, allows real-time mapping without external scanning mechanics. In , monochromators filter single-photon sources to suppress noise, enabling precise selection for entanglement experiments and .

References

  1. [1]
    Monochromators - RP Photonics
    A monochromator is a kind of tunable optical bandpass filter, which is in most cases tunable concerning the center wavelength of the pass band.
  2. [2]
    7.3: Wavelength Selectors
    ### Summary of Monochromator from https://chem.libretexts.org/Courses/Providence_College/CHM_331_Advanced_Analytical_Chemistry_1/07:_Components_of_Optical_Instruments_for_Molecular_Spectroscopy_in_the_UV_and_Visible/7.03:_Wavelength_Selectors
  3. [3]
    What is a Monochromator? - AZoOptics
    Aug 17, 2012 · A monochromator is an optical component which is used to produce a beam of light with a very narrow bandwidth - effectively light of a single colour.
  4. [4]
    7.3: Wavelength Selectors - Chemistry LibreTexts
    Feb 14, 2023 · A monochromator is an alternative method for selecting a narrow band of radiation that also allows us to continuously adjust the band's nominal ...
  5. [5]
    Spectrometers, Monochromators and Spectrographs - HORIBA
    A monochromator produces a beam of light with an extremely narrow bandwidth, or light of a single color. It is used in optical measuring instruments where ...
  6. [6]
    History of Spectroscopy | Ossila
    Spectroscopy In The 19th century​​ They used a slit rather than a rounded aperture to disperse the light. The first person to do so was W. H. Wollaston in 1802 ...
  7. [7]
    None
    Below is a merged and comprehensive summary of the historical timeline of spectrograph and monochromator development from the 1800s to the 1900s, combining all the information from the provided segments. To retain maximum detail in a dense and organized format, I will use a table in CSV format for key events, followed by a narrative summary of milestones, key inventors, and useful URLs. This approach ensures all data is preserved and easily accessible.
  8. [8]
    Landmark Publications in Analytical Atomic Spectrometry
    The landmark paper by Kirchhoff and Bunsen published in 1860 (English translation is available) is well known and was recommended by many scientists, including ...
  9. [9]
    Progress in the Preparation and Characterization of Convex Blazed ...
    Wood [5] invented the “blazed” grating technique in 1910 to increase the diffraction efficiency of the grating. This technique alters the energy distribution ...Missing: history | Show results with:history<|separator|>
  10. [10]
    Brief Company History - McPherson
    1968- Wisconsin University receives the first Schoeffel mini-double Czerny-Turner monochromator ... monitoring; 2005- ANKA (Karlsruhe) gets 27m beamline for ...Missing: date computer
  11. [11]
    Discover 50 years of Raman innovation by HORIBA
    The Lirinord HG2S double monochromator was the first Raman spectroscopy instrument equipped with concave holographic gratings ... The famous Jobin –Yvon U1000 in ...Missing: key milestones
  12. [12]
    50 years of Diffraction Gratings - HORIBA
    Monochromator ... 1969 First commercialization of flat holographic gratings. 1969 Patent of Jean Flamand, Antoine Labeyrie Diffraction Gratings [3,628,849].
  13. [13]
    Acousto-optic tunable filters - ADS - Astrophysics Data System
    This paper reviews the theory, techniques, and applications of acousto-optic tunable filters (AOTF). Two basic types of AOTF are described, i.e., ...
  14. [14]
    Fiber-to-fiber monochromator with high light throughput
    The high-quality fiber-to-fiber coupling of the monochromator allows flexible integration into all systems, while the design of the optical gratings and ...
  15. [15]
    [PDF] Chapter 7 Components of Optical Instruments
    Monochromators: • Entrance slit. • Collimating lens or mirror. • Dispersion element (prism or grating). • Focusing lens or mirror. • Exit slit. FIGURE 7-18 Two ...
  16. [16]
    The Monochromator and Its Role in the Spectrograph
    A monochromator is a device that separates different wavelengths of light from a given light source.
  17. [17]
    Monochromator: Fundamental Principle and Methods
    A monochromator is an optical device that converts polychromatic light (such as sunshine or light from a lamp) into a range of individual wavelengths ...
  18. [18]
    Monochromator System Optics - HORIBA
    In any spectrometer system, a light source should be imaged onto an entrance slit (aperture) which is then imaged onto the exit slit and so on, to the detector, ...
  19. [19]
    Calculating Monochromator Throughput - Newport
    The purpose of these two examples is to help you estimate the available monochromatic power or monochromator throughput for a detection system.
  20. [20]
    Diffraction Grating Physics - Newport
    A diffraction grating is a multi-slit surface that provides angular dispersion, separating wavelengths based on the angle they emerge from the grating.
  21. [21]
    [PDF] Section 18 Dispersing Prisms
    The details of the prism dispersion depend on the geometry used and the index dispersion ... Glass Dispersion 0 dn dλ. = <. 0 d d δ λ. > Blue light is deviated ...Missing: explanation | Show results with:explanation
  22. [22]
    [PDF] Chapter 5: Diffraction
    The grating equation:​​ The previous discussion showed that an incoming plane perpendicular to the grating results in one plane wave angled up at an angle “theta ...
  23. [23]
    The Grating Equation - Newport
    The grating equation mλ= d (sinα + sinβ) (2-1) which governs the angular locations of the principal intensity maxima when light of wavelength λ is diffracted ...
  24. [24]
    1.4: Spectroscopy and the Electromagnetic Spectrum
    Dec 27, 2021 · ... energy of a particular wavelength can be expressed as: (12.5.1) E = h ⁡ c λ. where E is energy in kcal/mol, λ (the Greek letter lambda) ...
  25. [25]
  26. [26]
    Dispersion (of Materials) - an overview | ScienceDirect Topics
    The wavelength dispersing properties of prism-based monochromators depend upon the material dispersion of the prism: the variation in the refractive index ...
  27. [27]
    theory and principles of monochromators, spectrometers and ...
    Aug 9, 2025 · Many of these instruments use prisms or gratings as dispersing components. There are a few basic types of these instruments utilizing the above ...
  28. [28]
    All About Diffraction Gratings
    ### Summary of Diffraction Gratings for Spectroscopy/Monochromators
  29. [29]
    Diffraction Gratings Tutorial - Thorlabs
    Sep 1, 2022 · Diffraction gratings, either transmissive or reflective, can separate different wavelengths of light using a repetitive structure embedded within the grating.
  30. [30]
    Diffraction Gratings - RP Photonics
    Echelle gratings are a special type of echelette gratings (= blazed gratings), where the blaze angle is particularly large (beyond 45°). They are typically made ...Details of Diffraction at a Grating · Blazed Gratings · Diffraction Efficiency
  31. [31]
    Diffraction Grating Selection Guide - Newport
    An R2 echelle, for example, has a blaze angle of tan–1(2) = 63.4°; an R5 echelle has a blaze angle of tan–1(5) = 78.7°. R number, Groove angle. R1, 45.0°. R2 ...Missing: lambda / | Show results with:lambda /
  32. [32]
    High-resolution, high-efficiency narrowband spectroscopy with an ...
    We describe and test a novel method for reaching a high spectral resolution with very high unpolarized diffraction efficiency in first-order2 Theory · 2.1 Vph Gratings · 3 Results
  33. [33]
    Echelle Grating Spectroscopic Technology for High-Resolution and ...
    Oct 31, 2022 · Echelle grating provides high spectral resolving power and diffraction efficiency in a broadband wavelength range by the Littrow mode.Echelle Grating... · 2.2. Blaze In Full Spectrum · 3. Optical Layout
  34. [34]
    Spectrometer Configurations: Littrow, Czerny-Turner and More | Ossila
    The Ebert-Fastie (sometimes referred to as Fastie-Ebert) configuration is composed of a large spherical mirror and a single plane diffraction grating. The light ...Optical Configurations · Ebert-Fastie Configuration · Czerny-Turner Configuration
  35. [35]
    Grating Monochromator Designs - Newport
    The Ebert mount. A single concave mirror replaces the two concave mirrors found in Czerny-Turner mounts. This design is a special case of a Czerny-Turner ...
  36. [36]
    [PDF] The Optics Of Spectroscopy
    2.1 Basic Designs. ○. 2.2 Fastie-Ebert Configuration. ○. 2.3 Czerny-Turner Configuration. ○. 2.4 Czerny-Turner/Fastie-Ebert PGS Aberrations. 2.4.1 Aberration ...
  37. [37]
    [PDF] Optics of Spectroscopy Tutorial
    The Czerny-Turner (CZ) monochromator consists of two concave mirrors and one ... gratings as well as Czerny - Turner and Fastie - Ebert spectrometers. Lambda ...
  38. [38]
    [PDF] Czerny Turner Monochromator - COMSOL
    This model simulates a grating spectrometer in a crossed Czerny-Turner configuration. The model uses the Geometrical Optics interface to compute the positions ...
  39. [39]
    1m High-throughput High-resolution Monochromator | McPherson
    Model 2061 Czerny-Turner Monochromator Higher Resolution and Throughput ... Resolution (nm) at 313.1-nm, 0.009, 0.012, 0.018, 0.036, 0.07, 0.14, 0.28, 1.05.Missing: commercial | Show results with:commercial
  40. [40]
  41. [41]
    [PDF] 9072, 9072da/ds monochromators - Sciencetech Inc.
    The double dispersion of 9072DA results in a better optical resolution compared to a single 9072 monochromator while the bandpass is half that of a single 9072.
  42. [42]
    Double Monochromators: Additive or Subtractive? - OLIS CLARiTY
    Two key attributes of the subtractive are very desirable: the production of a homogenous output beam and zero temporal dispersion.
  43. [43]
    Double Additive/Subtractive Monochromator - Catalog - Standa
    Optical schematic, Modified Czemy-Turner double monochromator supplemented with an optical matching unit providing either additive or subtractive dispersion.Missing: types | Show results with:types
  44. [44]
    [PDF] FLS1000 - Edinburgh Instruments
    Stray Light Rejection. 1:10-5 (single), 1:10-10 (double). Gratings. Mounted to triple grating turret. Accuracy. ± 0.2 nm *. Minimum Step Size. 0.01 nm *.
  45. [45]
    Gemini 180 Double monochromator - HORIBA
    The Gemini 180 is a fully automated, double additive grating scanning monochromator. The incorporation of toroidal optics provides for optimum throughput and ...
  46. [46]
    Resolution of Monochromators and Spectrographs - Newport
    The entrance slit width generally determines the spectral resolution of a spectrograph. However, the limiting resolution is reached when the entrance slit is ...
  47. [47]
    Understanding Spectral Resolution & Bandpass in Spectroscopy
    Aug 21, 2025 · The intensity of light passing through the monochromator is roughly proportional to the square of the slit width. Wider slits allow ...
  48. [48]
    Stray Light in Monochromators and Spectrographs - Newport
    For a major improvement in the signal to stray light ratio, use a double MS257 monochromator. ... To reduce stray light and increase throughput, the 77259 F/# ...Missing: bandwidth | Show results with:bandwidth
  49. [49]
  50. [50]
    Spectral Ghosts: Diffraction Grating Stray Light
    Jun 25, 2025 · Spectral ghosts are a stray light phenomenon caused by repeating, periodic imperfections on a ruled diffraction grating of an optical instrument.
  51. [51]
    [PDF] Overview of Stray Light Part 1
    • Proper baffle design minimizes vignetting and improves rejection of scattered light. • There will always be some vignetting if baffles are used. • Paths may ...Missing: monochromator double stages
  52. [52]
    [PDF] A Double Monochromator on a Spectrofluorometer - HORIBA
    A double monochromator reduces stray light by having two dispersive grating elements, one after the other, multiplying stray light rejection.
  53. [53]
    BAuA - Optical Radiation - Optical measurement technology - BAuA
    While a single monochromator is characterised by a stray light rejection of 10-4 (corresponds to a signal-to-noise ratio of 104), it is significantly better ...
  54. [54]
    What is stray light? - Shimadzu Scientific Instruments
    In theory stray light dictates the dynamic range for most spectrophotometers. In general, the lower the stray light specification, the more expensive the ...
  55. [55]
    Ultraviolet Detectors: Perspectives, Principles, and Practices
    Oct 1, 2019 · At absorbance above 2, very low light intensity is transmitted in the sample beam, so the amount of stray light (background light detected) ...
  56. [56]
    Monochromators: How Do They Work? - Ossila
    Adjustable Slits: The components that control how much light enters and leaves the monochromator. · Concave Mirrors: Used to focus, redirect and collimate light.Missing: path | Show results with:path
  57. [57]
    Materials Data - Fused Silica / Fused Quartz - Esco Optics, Inc.
    Ultraviolet (UV) Fused Silica. Optimum Transmission Range: Ideal Applications: 180 nm - 2.0 µm. Deep UV transmission elements. Minimum scattering systems. UV ...Missing: monochromator | Show results with:monochromator
  58. [58]
    How to Specify a Monochromator - McPherson Inc.
    Stray Light Rejection If low stray light is important add aggressive filtering to your setup. You might also use a double monochromator. This doubles the ...Missing: tilted | Show results with:tilted
  59. [59]
    Calcium Fluoride Windows - Thorlabs
    Uncoated CaF 2 Wavelength Range: UV to IR (180 nm - 8 μm); Available AR Coated for 1.65 - 3.0 µm (Ravg < 1.0%). Thorlabs offers calcium fluoride (CaF 2) High ...
  60. [60]
    7.3: Wavelength Selectors
    ### Summary on Prism Dispersion and Comparison to Gratings
  61. [61]
    Dispersion - Newport
    D = dβ/dλ = m / dcosβ = (m/d)secβ = Gm secβ (2-13). where β is given by Eq. (2-4). The quantity D is called the angular dispersion. As the groove frequency G ...Missing: dθ/ cos θ)
  62. [62]
    [PDF] Choosing a Monochromator or Spectrograph for Chemical Analysis
    But the trade-off is speed: It can take several minutes to scan a large spectral range with high spectral resolution in mono- chromator mode. Smaller ...
  63. [63]
    [PDF] Soft X-ray Monochromators - Luxium Solutions
    Crystals of the acid phthalate family offer the wide interplanar spacing required for diffracting soft x-rays and permit spectroscopy and imaging from 2.3 to 25 ...
  64. [64]
    [PDF] Multilayer monochromators for EUV and X-ray optics and the ... - HAL
    Mar 23, 2015 · However, the wavelength ranges are determined by the lattice spacing of the available natural crystals, which are too small to reflect EUV and ...
  65. [65]
    Echelle Gratings - an overview | ScienceDirect Topics
    An echelle grating is a coarsely ruled grating with a high blaze angle, allowing higher diffraction orders for increased dispersion and resolution.
  66. [66]
    Aspheric Ebert Monochromator for Simultaneous Observation of ...
    Thus, after correcting coma, we are left with spherical aberration which could be corrected entirely by the asphericity on the plate.
  67. [67]
    Getting Light into a Monochromator - Newport
    The slit width is usually selected to achieve the bandpass or spectral resolution required. To optimize throughput, always use the widest slits you can, while ...Missing: Δλ = | Show results with:Δλ =
  68. [68]
    Anamorphic Magnification - Newport
    The ratio b/a is called the anamorphic magnification. For a given wavelength λ, it depends only on the angular configuration in which the grating is used.
  69. [69]
    Spectrometer Throughput and Etendue - HORIBA
    Because flux, etendue, and radiance must be conserved between object and image ... The etendue of the spectrometer is given by Equation (52). Then ...
  70. [70]
    [PDF] Modeling of Spectrometer using Zemax OpticStudio Simulation
    A spectrometer is modeled using Zemax OpticStudio, a software for optical design and simulation, to evaluate performance and simulate light propagation.
  71. [71]
    FEA-based heat management of a monochromator at the High ...
    Aug 8, 2019 · The finite element method (FEA) was used for thermal mechanical behavior analysis of a monochromator subjected to high heat load for a ...
  72. [72]
    Bandpass and Resolution - HORIBA
    Both bandpass and resolution are used as characterization of an instrument's ability to separate adjacent spectral lines.
  73. [73]
    [PDF] Monochromators - Laser Components
    As shown in the optical diagram, polychromatic radiation is focused at the entrance slit and reflected by a folding mirror onto a spherical collimating/.<|separator|>
  74. [74]
    Understanding slit width, grating, and optical principles in ...
    Jun 19, 2025 · You use a grating as the main dispersive element in a monochromator. When collimated light hits the grating, it creates a diffraction ...
  75. [75]
    Beer-Lambert Law and Visible Light Spectrometers - Harper College
    As the Beer-Lambert Law does not hold for multiple wavelengths, a monochromator (historically this contained a prism, but currently gratings are used) is used ...<|separator|>
  76. [76]
    Monochromators - Shimadzu
    The monochromator comprises a dispersive element, an entrance slit and mirrors to create a parallel beam similar to sunlight, and an exit slit and mirrors to ...
  77. [77]
    [PDF] A Double Monochromator on a Spectrofluorometer - HORIBA
    Sep 17, 2019 · A double monochromator reduces stray light by having two dispersive grating elements, multiplying the stray light rejection of a single  ...
  78. [78]
    Premonochromator for stray light rejection in Raman spectroscopy
    Replacing the ruled gratings in the double monochromator with holographic gratings considerably improves the stray light rejection (10−14 at 20 cm−1 from the ...Missing: tilted | Show results with:tilted
  79. [79]
    Raman Spectroscopy - Omega Optical
    It's possible to use a double or triple grating spectrometer to accomplish rejection of the background signal.
  80. [80]
    Photomodulated infrared spectroscopy by a step-scan Fourier ...
    Nov 3, 2006 · A conventional PR experimental system is based on a dispersive monochromator. Generally, it consists of a monochromator, a combination of pump ( ...
  81. [81]
    Wavelength modulation spectroscopy using novel mechanical light ...
    Jul 21, 2005 · We describe two mechanical light chopper blade designs that can be used, with a monochromator and broadband light-source-based spectroscopy ...
  82. [82]
    Grating monochromators for the spectral selection of femtosecond ...
    Time-preserving monochromators are realized in a single-grating configuration if the number of illuminated grooves is minimum for a given spectral resolution; ...
  83. [83]
    Advanced time-resolved absorption spectroscopy with an ultrashort ...
    May 11, 2021 · Ultrashort visible-near infrared (NIR) pulse generation and its applications to ultrafast spectroscopy are discussed.
  84. [84]
    [PDF] Towards wavefront preservation X-ray crystal monochromator for ...
    To illustrate the consequences of the thermal deformation of the crystal on the beam intensity distribution, we have focused on the crystal C1-1 – Si(111) with ...
  85. [85]
    [PDF] X-ray monochromators - Cheiron School 2010
    Energy, energy resolution. Spatial spread. Time structure. Angular divergence intensity. Controlling the X-ray beam properties by X-ray crystal monochromators ...
  86. [86]
    Short Wavelength X-Rays - an overview | ScienceDirect Topics
    For diffraction applications, only short-wavelength X-rays in the range of a few angstroms to 0.1 Å, which means in the range of 1–120 keV energy, are used.
  87. [87]
    Italian panoramic monochromator for the THEMIS telescope
    The Panoramic Monochromator substantially is a narrow band filter (approximately equals 22 mAngstrom bandwidth) tunable on the visible spectrum for quasi ...
  88. [88]
    Fabry–Perot Monochromator - Oxford Academic
    Abstract. Many astronomical spectrographs are inefficient in the sense that light from the telescope image is both rejected at the entrance slit and wasted.Missing: spectroscopy | Show results with:spectroscopy
  89. [89]
    Laser Tuning With Diffraction Gratings - Newport
    The combination of a diffraction grating and one of these dyes in a laser allows the production of laser light that has a narrow line width and which is tunable ...
  90. [90]
    Tunable OPO Lasers: Q-Tune Series for Spectroscopy & Beyond
    Jul 10, 2025 · Tunable OPO lasers, like the Q-TUNE series, use an OPO to generate tunable light, with the Q-TUNE series having a range of 410-2300 nm.
  91. [91]
    High efficiency diffraction grating for EUV lithography beamline ...
    A high efficiency blazed grating for use in an EUV lithography beamline monochromator was successfully fabricated and tested. ... EUV lithography beamline ...
  92. [92]
    Design and analysis of multi-color confocal microscopy with a ...
    May 10, 2012 · In this paper, confocal-spectral microscopy is proposed with a particular spectrometer design with a wavelength-scanning galvano-mirror. This ...Missing: monochromators | Show results with:monochromators
  93. [93]
    MEMS tunable grating micro-spectrometer - SPIE Digital Library
    Nov 21, 2017 · The proposed concept is based on a monochromator realized using tunable MEMS blazed gratings. The MEMS grating is stretched like an accordion, ...
  94. [94]
    Monochromator Selection Guide - Bentham Instruments
    The Bentham Monochromator Selection Guide has been written to allow new users to quickly and easily identify the most appropriate monochromator configuration.Missing: limitations | Show results with:limitations
  95. [95]
    A Convenient Technique for Determining the Relative Quantum ...
    A technique is described for determining the relative efficiency with wavelength of a monochromator and detector system using single photon detection.