Fact-checked by Grok 2 weeks ago

White light interferometry

White light interferometry (WLI) is a non-contact optical measurement technique that employs , incoherent light sources, such as white light, to produce patterns characterized by short lengths, allowing for high-precision determination of surface topography, thin-film thickness, and other without the multiple fringe ambiguities common in monochromatic . The method relies on the principle that constructive occurs only when the difference between reference and sample beams is near zero, typically within a few micrometers, enabling unambiguous localization of the envelope through vertical scanning. The origins of WLI trace back to the 17th century with Isaac Newton's early observations of using white light, though systematic development began in the when employed it in 1893 to measure the of gases and etalon thicknesses. Key advancements occurred in the mid-20th century, including James M. Burch's 1953 invention of the scatterplate interferometer for path matching and André Henri Mirau's 1952 patent of the Mirau interferometer objective, which facilitated microscopic surface profiling. The and saw significant refinements with the integration of low-coherence sources for characterization and biomedical imaging, culminating in the 1990s commercialization of scanning white-light interferometers for industrial . In practice, WLI systems often incorporate a Mirau or configuration within a microscope objective, where broadband illumination (typically 400–700 nm) from a halogen lamp or LED passes through beam splitters to create superimposed reference and object wavefronts. The resulting interferogram exhibits a central high-contrast fringe packet whose peak position is detected via phase-shifting or envelope correlation algorithms, yielding vertical resolutions down to 1 nm and lateral resolutions approaching 0.5 μm, limited by the numerical aperture. Advantages include robustness to environmental vibrations through short measurement times (milliseconds) and the ability to profile discontinuous or rough surfaces with step heights up to hundreds of micrometers, as the coherence gate eliminates spurious fringes. WLI finds broad applications in precision manufacturing, such as characterizing micro-electro-mechanical systems () and surfaces, where it quantifies roughness and form errors with sub-nanometer accuracy. In , it assesses surface modifications from processes like ion sputtering (e.g., crater depths on at rates of 1–10 nm per pulse) and , enabling precise volume loss calculations in studies. Biomedical extensions, including low-coherence variants for (OCT), support retinal imaging and tissue characterization with axial resolutions of 10–15 μm. Recent developments emphasize compensation for in-situ measurements and enhanced resolution via UV illumination or super-resolution techniques, expanding its utility in dynamic environments.

Principles

Basic Principles

White light interferometry is a non-contact optical technique for measuring surface heights and profiles using incoherent sources. This method relies on the interference of reflected from a reference surface and the sample under test to generate high-resolution topographical data. Interference fringes form through the superposition of split into reference and object paths, which are recombined after . Due to the short of the broadband source, fringe visibility is confined to a narrow region near zero difference, enabling precise localization of the signal. White light interferometry originated from adaptations of the developed in the late 19th century for precise length measurements, with significant advancements in the integrating modern and computing for surface profiling applications. A primary advantage is the provision of ambiguity-free absolute distance measurements over extended ranges, unlike monochromatic laser interferometry, where phase ambiguity limits the measurement depth; this stems from the envelope that envelopes the fringes. The quantifies this fringe localization, as detailed in subsequent sections on concepts. The intensity of the interferogram as a function of difference \delta is described by I(\delta) = I_0 \left[1 + V(\delta) \cos\left(\frac{2\pi \delta}{\lambda}\right)\right], where I_0 is the average background intensity, V(\delta) is the visibility function that decays with increasing \delta, and \lambda is the central wavelength of the source.

Coherence Concepts

In white light interferometry, the temporal coherence length l_c is defined as the optical path difference over which the phase correlation between light waves from a broadband source remains predictable, typically on the order of microns for white light sources. The coherence length is approximately given by l_c \approx \frac{\lambda_0^2}{\Delta \lambda}, where \lambda_0 is the central wavelength and \Delta \lambda is the spectral bandwidth (FWHM). This short l_c arises because the broad spectral content of white light leads to rapid dephasing of interfering waves as the path difference increases, limiting observable interference to a narrow range around zero optical path difference (OPD). In contrast to monochromatic laser sources with long coherence lengths, white light's limited l_c ensures that interference fringes are confined to a specific axial position, facilitating precise surface profiling without ambiguity. The envelope describes the Gaussian-like intensity packet that modulates the fringes, within which the fringes exhibit high and are centered at zero OPD. Outside this envelope, the fringes fade rapidly due to destructive superposition across the source's wavelength spectrum, effectively localizing the signal to a small volume. This envelope's width is directly tied to l_c, providing a gate for axial in interferometric setups. The short l_c in white light interferometry enables absolute positioning of surfaces by scanning the reference arm until the point of maximum fringe contrast, which corresponds to the zero OPD and avoids the $2\pi ambiguity inherent in monochromatic phase-shifting methods. This approach yields unambiguous measurements over ranges exceeding typical fringe spacings, making it suitable for discontinuous or rough surfaces. Factors influencing include the source's bandwidth, which inversely affects l_c, the central that scales the sensitivity, and distinctions between temporal (governed by ) and spatial (dependent on source size and beam collimation). The visibility function V(\delta), which quantifies as a function of OPD \delta, decays qualitatively from unity at \delta = 0 to near zero beyond l_c, forming the envelope's shape. For example, spectra from LED or sources, with bandwidths around 20-100 nm centered near 600 nm, typically yield l_c \approx 2-15 \, \mu \mathrm{m}, allowing sub-micron axial localization in practical systems. This decay ensures that only path-matched regions contribute to the signal, enhancing measurement specificity.

Instrumentation

Interferometer Setups

White light interferometry commonly employs the configuration for measuring displacements and surface profiles, where a divides broadband light into two arms: a reference arm with a fixed mirror and a sample arm directed toward the test surface. The recombined beams interfere only when their lengths are closely matched, leveraging the short to localize the fringe packet. A variant, the Twyman-Green interferometer, is adapted for precise surface testing by using a collimated beam and a high-quality transmission flat or reference mirror in the reference arm, while the test arm accommodates the surface under inspection, such as a lens or mirror. This setup allows for the generation of interference patterns that reveal wavefront deviations, with adjustments via lenses to match curvatures of non-planar test objects. Fiber-optic implementations enable applications by integrating a coupler to split the between a reference and a sensing , with reflections recombined and detected at the output for compact, flexible configurations. These systems often incorporate GRIN lenses and actuators for path modulation, suitable for in-situ measurements in harsh environments. Broadband light sources are essential, typically tungsten-halogen lamps providing a continuous from visible to near-infrared wavelengths or superluminescent diodes (SLDs) with spectral widths exceeding 50 nm to ensure short lengths under 100 µm. These sources deliver sufficient intensity for while maintaining the low-coherence properties critical for determination. Detector arrays, such as or sensors, capture two-dimensional interferograms in imaging setups by recording the spatial variation of interference intensity across the field of view. These arrays enable high-resolution mapping of surface through pixel-wise analysis of the fringe envelope. Alignment in these setups requires minimizing group delay in optical components to avoid broadening the envelope, and controlling mismatches to prevent visibility loss in the interferogram. Achromatic beam splitters and polarization-maintaining fibers are often used to preserve fringe contrast.

Microscopic Implementations

White light interferometry (WLI) adaptations for microscopic applications integrate specialized objectives and scanning mechanisms to enable high-resolution surface at the microscale, typically for rough or discontinuous surfaces with sub-micrometer lateral and nanometer vertical precision. These implementations leverage compact interferometer designs within systems to combine illumination with detection, allowing non-contact measurements over small fields of view. Mirau objectives employ an infinity-corrected design where a and reference mirror are integrated directly into the objective lens assembly, facilitating compact epi-illumination setups for magnifications up to 50x without requiring separate interferometer components. This configuration splits the incident light into reference and sample paths within the objective, recombining them to form fringes at the detector, which is particularly advantageous for vibration-sensitive environments due to its short differences. In contrast, Linnik objectives use a similar infinity-corrected approach but incorporate two matched high-numerical-aperture objectives—one for the sample and one for the reference arm—to achieve higher magnifications, such as 100x or more, while maintaining of the beams for precise alignment. The dual-objective design in Linnik setups minimizes aberrations but requires careful matching of the to ensure uniform across the field. Vertical scanning interferometers in microscopic WLI systems typically feature a piezoelectric (PZT)-driven sample stage that provides precise axial translation, enabling the sample to scan through the focal plane to locate the envelope for absolute height measurements. The PZT stage offers sub-nanometer step over scan ranges up to several millimeters, allowing the system to capture envelope peaks from broadband sources with short lengths for unambiguous topography reconstruction. Commercial systems exemplify these implementations, such as the Zygo NewView series, which utilizes scanning interferometry with Mirau or Linnik objectives to achieve vertical resolutions below 0.1 nm and fields of view up to several millimeters for 3D surface mapping. Similarly, the ContourGT employs vertical scanning WLI with integrated objectives, delivering sub-nanometer Z-resolution over millimeter-scale areas through automated stage control and broadband LED illumination. Illumination in these microscopic setups often incorporates Köhler configuration to ensure uniform delivery of broadband white light across the sample, minimizing intensity variations and optimizing the for fringe contrast. This involves imaging the source at the condenser aperture while focusing a uniform field at the sample plane, which is essential for maintaining consistent properties over the imaged area. A key challenge in microscopic WLI is managing chromatic dispersion within the objectives, which can broaden the coherence envelope and degrade axial resolution due to wavelength-dependent path lengths. This is addressed through the use of low-dispersion glasses or achromatic designs that minimize group delay variations across the , ensuring sharp localization of the peak. Typical resolution limits in these systems include lateral resolutions around 0.5 μm, determined by the objective's and limits, and vertical resolutions approaching 1 , enabling detailed inspection of microstructures in microelectromechanical systems () and devices. Such capabilities support non-destructive profiling of features like trenches and steps in fabrication, where sub-micrometer variations critically affect performance.

Operation

Operating Modes

White light interferometry primarily operates in vertical scanning mode, where the sample or objective lens undergoes axial translation to systematically vary the difference (OPD) between the reference and measurement arms of the interferometer. This scanning process generates a series of interferograms, or correlograms, from which the position of the envelope maximum is identified to map surface heights, as the short of white light localizes high-contrast fringes to near-zero OPD. To achieve sub-fringe resolution, phase-shifting interferometry (PSI) is integrated into the vertical scanning mode, typically in a five-frame sequence, while white light scanning provides coarse positioning over larger height ranges. This hybrid approach combines the high precision of PSI (on the order of nanometers) with the extended measurement range enabled by the coherence envelope for unambiguous height determination. Coherence peak tracking enhances real-time operation by continuously monitoring the modulation contrast or fringe visibility during the scan, allowing identification of the envelope maximum without post-processing the full dataset. This method facilitates dynamic adjustment of the scan and is particularly useful for surfaces with varying topography. Data acquisition in these modes involves capturing sequential interferogram frames at fixed axial intervals, such as 0.2 μm steps, over a range typically spanning several times the source's (e.g., 10–100 μm depending on ) to ensure the is fully resolved. Environmental controls are essential to maintain accuracy, including tables and temperature stabilization to minimize effects on the . Mode selection depends on the application: vertical scanning is standard for detailed measurement of localized features, while lateral scanning—where the sample or reference is translated horizontally with a tilted plane—enables efficient profiling of extended fields without stitching in one , suitable for larger or elongated surfaces.

Data Analysis Techniques

In white light interferometry (WLI), data analysis begins with processing the captured interferograms, typically obtained through vertical scanning modes where the sample is moved relative to the focal plane to generate a of images I(δ) as a function of difference δ. The primary goal is to extract surface height information z from the envelope and within each pixel's interferogram. This involves identifying the position of the zero-order , where the peak occurs, to determine the coarse height, followed by fine adjustments using data. Algorithms must handle the limited of white light sources, typically on the order of micrometers, to achieve sub-nanometer vertical resolution over ranges up to millimeters. Envelope detection locates the coherence peak δ_max, which approximates the surface height z ≈ δ_max / 2 due to the round-trip path in the interferometer. Common methods include the , which computes the to yield the as |I(δ) + i \mathcal{H}{I(δ)}|, where \mathcal{H} denotes the , enabling precise of the interferogram. Alternatively, intensity modulation analysis derives the from the squared modulus of the complex interferogram, often using nonlinear algorithms adapted from phase-shifting techniques for computational efficiency. These approaches are robust to noise and provide the envelope shape, essential for peak localization in low- signals. To compute the envelope maximum with sub-pixel accuracy, techniques such as the centroid method or Gaussian fitting are employed. The centroid method calculates the peak position as z = \frac{\int \delta \, I(\delta) \, d\delta}{\int I(\delta) \, d\delta}, weighting the intensity distribution around the window to achieve precisions below 1 , particularly effective for asymmetric envelopes caused by . Gaussian fitting models the envelope as I(δ) ≈ A \exp\left[-(δ - μ)^2 / (2σ^2)\right], where μ provides the peak location, offering high accuracy for well-defined peaks but requiring more computation. Phase extraction refines the height measurement within the narrow coherence window (±σ around δ_max), leveraging the monochromatic-like fringes there. Fourier transform methods apply a fast Fourier transform (FFT) to the interferogram to isolate the carrier frequency and unwrap the phase φ(δ), yielding the fine height correction Δz = \frac{\lambda}{4\pi} \Delta\phi, where λ is the central wavelength and Δφ is the phase difference. Phase-shifting interferometry (PSI) algorithms, such as those using multiple frames at the coherence peak, compute the phase via least-squares fitting of intensity equations, achieving resolutions down to λ/1000 without frequency-domain processing. For 3D surface reconstruction, height maps are generated pixel-wise by applying the above analyses to each interferogram in the scanned , producing a z(x,y) with lateral determined by the objective . Large-area measurements require stitching adjacent fields of view, using feature-based algorithms to align and blend overlaps while minimizing seam artifacts from tilt or . This enables extended scans over centimeters with maintained accuracy. Error sources in include speckle noise, arising from rough surfaces coherent components within the broadband source, which broadens the and reduces peak sharpness; mitigation involves averaging multiple acquisitions per position to suppress random fluctuations. Dispersion mismatches between interferometer arms distort the symmetry, leading to biased height estimates, and are corrected via chromatic optimization of or post-processing adjustments to the group delay. like MountainsMap implements these techniques, including automated detection, phase unwrapping, and profile generation compliant with ISO standards for .

Theory

Spectral Width and Coherence Length

In white light interferometry, the l_c represents the difference over which fringes maintain high , directly influencing the technique's ability to resolve surface heights unambiguously. This length is inversely proportional to the spectral bandwidth of the source, with the general approximation given by l_c \approx \frac{\lambda_0^2}{\Delta \lambda}, where \lambda_0 is the central wavelength and \Delta \lambda is the (FWHM) of the spectral width. This relation arises from the temporal properties of , enabling short l_c values that localize the zero-order fringe for precise measurements. The derivation of this relationship stems from the Fourier transform of the power spectral density G(\nu) of the light source, where the degree of coherence \gamma(\tau) is defined as \gamma(\tau) = \frac{\int G(\nu) e^{i 2\pi \nu \tau} \, d\nu}{\int G(\nu) \, d\nu}. The coherence time \tau_c corresponds to the delay \tau at which |\gamma(\tau)|^2 drops significantly (e.g., to $1/e), and l_c = c \tau_c, with the effective width determined by evaluating the denominator relative to the squared magnitude of the numerator at \tau = l_c / c. For a rectangular spectrum, this yields the approximate formula above, as the transform produces a sinc-like envelope whose width scales with $1/\Delta \nu, and \Delta \nu \approx c \Delta \lambda / \lambda_0^2. The choice of light source type significantly affects \Delta \lambda and thus l_c. For instance, white light-emitting diodes (LEDs) typically exhibit a broader \Delta \lambda of around 100 nm in the visible range, resulting in a shorter l_c (on the order of a few micrometers), which enhances for high-aspect-ratio surfaces such as deep trenches or stepped structures by confining to localized path matches and reducing in height . In contrast, narrower-band sources like filtered yield longer l_c, suitable for smoother profiles but prone to phase ambiguities over larger ranges. To illustrate, for visible light with \lambda_0 = 550 nm and \Delta \lambda = 50 nm, the approximation gives l_c \approx 6 \, \mu \mathrm{m}, providing a practical scale for micrometer-level surface profiling without extensive scanning. Units are consistent in vacuum or air ( n \approx 1), though in media requires adjustments by dividing by n. This approximation holds under the assumption of a rectangular spectrum; real sources, such as LEDs with Gaussian or irregular profiles, necessitate shape-specific corrections to the integral for accurate l_c estimation, as deviations can alter the envelope decay rate. Such adjustments ensure reliable application in interferometric setups.

Gaussian Spectrum Analysis

In white light interferometry, the Gaussian spectral shape is a prevalent model for broadband sources, characterized by the power spectral density G(\lambda) = \exp\left[ -(\lambda - \lambda_0)^2 / (2 \sigma_\lambda^2) \right], where \lambda_0 is the central wavelength and \sigma_\lambda is the standard deviation. The full width at half maximum (FWHM) spectral bandwidth is given by \Delta \lambda = 2 \sqrt{2 \ln 2} \, \sigma_\lambda \approx 2.355 \sigma_\lambda. The coherence properties arise from the Fourier transform relationship between the spectrum and the autocorrelation function, as per the Wiener-Khinchin theorem, yielding a Gaussian envelope for the interferogram visibility. Specifically, the degree of coherence, or envelope function V(\delta), for path length difference \delta is V(\delta) = \exp\left[ -\left( \frac{\pi \Delta \nu \delta}{2 \sqrt{\ln 2} \, c} \right)^2 \right], where \Delta \nu = c \Delta \lambda / \lambda_0^2 is the FWHM in frequency domain and c is the speed of light. This form reflects the inverse proportionality of the envelope width to the spectral standard deviation. The exact coherence length l_c, defined as the path difference where the envelope decays to $1/e of its maximum, is l_c = \frac{2 \sqrt{\ln 2}}{\pi} \frac{\lambda_0^2}{\Delta \lambda}, which is slightly shorter than the rectangular spectrum approximation l_c \approx \lambda_0^2 / \Delta \lambda. This expression refines the general relation between spectral width and coherence length by accounting for the Gaussian tails. The Gaussian model is particularly suitable for (LED) sources commonly used in white light interferometry, as their emission spectra often approximate this shape, allowing precise predictions of fringe visibility decay and interferogram localization without significant distortion from spectral irregularities.

Applications and Advances

Key Applications

White light interferometry (WLI) is widely employed in for non-contact topography measurements, enabling precise characterization of parameters such as Ra and Rz, as well as form errors in and semiconductors with nanometer-scale . This technique achieves sub-nanometer vertical precision, making it ideal for assessing smooth and rough surfaces in . Typical measurement volumes reach up to 10 mm × 10 mm × 1 mm, with mapping speeds around 1 second per scan, supporting high-throughput inspections. In optical testing, WLI excels at evaluating lens figure errors, mirror flatness, and aspheric surfaces, often in conjunction with null to compensate for surface deviations and ensure high-fidelity wavefront analysis. It provides nanometer accuracy for mid-spatial frequency errors on optical components, which are critical for performance in systems and . For and , WLI facilitates profiling of wafers, thin films, and nanostructures following processes, measuring parameters like step heights, film thicknesses, and with sub-nanometer resolution. This non-destructive approach supports in-line in production, enabling rapid assessment of device features down to the nanoscale. Biomedical applications of WLI include non-contact topography mapping of skin and tissue surfaces, as well as evaluation of dental crown fitting and implant surfaces for roughness and wear. In dermatology, it visualizes skin morphology in vivo without invasion, while in dentistry, it quantifies enamel etching and prosthetic fit with high precision. WLI also serves in displacement sensing for absolute position tracking in machine tools and short-range vibration analysis, offering sub-nanometer resolution over offsets up to several millimeters. This capability ensures stable measurements in dynamic environments, with active compensation for vibrations to maintain accuracy.

Recent Developments

Since the 2010s, white light interferometry (WLI) has seen significant advancements in acquisition speed, driven by techniques such as sub-Nyquist sampling and lateral scanning integrated with phase-shifting algorithms. In 2023, researchers introduced a deep learning-based sub-Nyquist sampling method for vertical scanning WLI, enabling envelope extraction from undersampled interferograms and achieving reconstruction speeds up to 10 times faster than traditional full-sampling approaches while maintaining sub-nanometer vertical resolution. Complementing this, a 2024 lateral scanning WLI system utilizing phase-shifting with fringe spacing tuned to multiples of 4 pixels per frame reached 80 frames per second using a standard camera, allowing stage speeds of 0.704 mm/s over a 5.25 mm × 1.25 mm —reducing measurement times by a factor of 4 compared to conventional methods without sacrificing 16.5 nm vertical . These enhancements have enabled mapping exceeding 100 in optimized setups, particularly for dynamic surface inspections. Accuracy improvements have focused on hybrid approaches combining WLI with phase-shifting interferometry () to achieve picoscale resolutions. A 2022 ultrathin WLI design, using division between two 140 μm glass plates, demonstrated resolutions of approximately 300 through self-calibration and tunable rotation, offering stability for absolute measurements with visibility up to 0.8 and negligible effects. Hybrid -WLI systems, such as phase-shifting configurations, further enhance precision by enabling simultaneous multi-phase acquisition, yielding sub-0.1 repeatability in transmissive modes for thin-film profiling. Miniaturization efforts have produced all-fiber and chip-scale WLI devices suitable for portable high-temperature sensing. Recent fiber-optic interferometric sensors, including sapphire-based extrinsic Fabry-Perot interferometers interrogated with white light, operate reliably up to 1000°C for simultaneous and , with resolutions better than 1°C in harsh environments like monitoring. These compact designs leverage sources for absolute positioning, facilitating integration into portable probes without cooling requirements. Post-2020 integrations of have improved WLI robustness via for noise suppression and automated analysis. A 2025 simulation-driven deep () approach reconstructs topography from noisy interferograms with root-mean-square errors below 47 nm—9% more accurate than methods—and processes 200 frames in under 0.4 seconds, enabling feature extraction on complex surfaces with up to 10% intensity noise. Hybrid WLI systems have expanded measurement ranges through combinations with complementary techniques. Bruker's 2023 ContourX-1000 and NPFLEX-1000 profilers incorporate Universal Scanning Interferometry, blending vertical scanning with sureface-finding algorithms for extended-depth measurements on large parts, achieving full analysis in 10 seconds with sub-nanometer precision across diverse materials. Looking ahead, on-chip WLI implementations promise broader adoption in inspection, enhanced by quantum techniques. Integrated photonic platforms using entangled photons for white-light have shown potential for surpassing classical limits in phase sensing, with ongoing research targeting scalable quantum for portable devices and telecom-integrated .

References

  1. [1]
    [PDF] White Light Interferometry
    For an interferometer to be a true white light achromatic interferometer two conditions need to be satisfied. First, the position of the zero order interference ...Missing: definition | Show results with:definition
  2. [2]
    Ultrathin picoscale white light interferometer - PMC - PubMed Central
    White light interferometers are key tools for non-invasive and non-contact measurement of surface topography via vertical scanning interferometry, thin-film ...
  3. [3]
    [PDF] An insight into optical metrology in manufacturing
    Feb 16, 2021 · In section 4, we introduce the basic principles and recent development trends of scanning white-light interfero- metry (SWLI) and ...
  4. [4]
    Characterization of Surface Modifications by White Light Interferometry
    Feb 27, 2013 · In this article, a number of applications of white light interferometry for investigation of surface modifications in mass spectrometry, and wear phenomena in ...
  5. [5]
    The Development, Commercialization, and Impact of Optical ...
    Low-coherence or white-light interferometry was first described by Sir Isaac Newton and in the 1980s it was used in optical communications to characterize ...
  6. [6]
    Principles of interference microscopy for the measurement of surface ...
    Interference microscopy plays a central role in noncontact strategies for process development and quality control, providing full 3D measurement of surface ...
  7. [7]
    Surface recovery algorithm in white light interferometry based on ...
    From the fundamental WLI principle, the light reflected from the surface forms interference fringes, and the period of one fringe is λ / 2 of the surface ...
  8. [8]
    Compensation of phase change on reflection in white-light ...
    White-light interferometry produces short coherence interferograms whose fringe visibility is narrowly localized such that the optical path difference between ...
  9. [9]
    Theoretical measurement uncertainty of white-light interferometry on ...
    White-Light Correlogram. The light intensity at the output of the Michelson interferometer as a function of the path difference z o - z r has the following form.
  10. [10]
    White Light Interferometers - RP Photonics
    White light interferometers are interferometers using broadband light inputs. They can be used for measuring chromatic dispersion of optical components, ...
  11. [11]
    Principles of white-light interferometry to Michelson - Polytec GmbH
    Whitelight-interferometry is the only measurement method where the resolution is independent of the field-of-view.
  12. [12]
    Ultrathin picoscale white light interferometer | Scientific Reports
    May 23, 2022 · One can also calculate the corresponding coherence time of the source as t_c = L_c/c, where c is the speed of light in vacuum. For the red LED ( ...<|separator|>
  13. [13]
    Coherence properties of different light sources and their effect on the ...
    Jul 19, 2017 · Temporal coherence is related to the intrinsic spectrum bandwidth of the light source, while spatial coherence can be affected by the size of ...
  14. [14]
    Metrology for Laser Optics
    ### Summary of Michelson Interferometer Setups in White Light Interferometry
  15. [15]
  16. [16]
    Twyman–Green Interferometers – optical surface inspection
    Figure 1: Setups of Twyman–Green interferometers. In the lower case, the wavefront curvature is adjusted to the curved test object, using an additional lens.
  17. [17]
    [PDF] Chapter 3: White-Light Michelson Interferometry with Optical Fibers
    Jan 3, 2015 · Simply described, a Michelson interferometer splits an incident beam into two parts, reflects each of the beams off of a different reflector, ...Missing: history | Show results with:history
  18. [18]
    [PDF] Interferometry with broadband light: Applications in metrology
    In a broadband light source such, as a filament lamp (white light) or Super. Luminescent Diode (SLD), a range of wavelengths are simultaneously present.
  19. [19]
    White Light Interferometers(WLI) | KEYENCE America
    Light interference occurs when two waves of light converge, causing the amplitude of the now combined wave to increase or decrease.
  20. [20]
  21. [21]
    Polarizing white light interferometry for phase measurements using ...
    Sep 15, 2023 · This study proposes a new design for a polarization phase-shifting white light interferometric system. The optical setup involves a polarizing ...
  22. [22]
    A Primer on White Light Interferometry and White ... - Evident Scientific
    In this white paper, we explain the technical details behind our novel gradient contrast method.
  23. [23]
    Dispersion optimized white-light interferometer based on a ...
    May 27, 2011 · Chromatic aberrations lead to systematic measuring errors in SWLI, especially on micro-structures with curved or tilted surface areas. For ...
  24. [24]
    [PDF] Construction and Validation of a White Light Interferometer
    3.3 White-Light Interferometry Equation. In the white-light interferometry, the detected light intensity I is a function of the reference mirror scanning ...
  25. [25]
    Dispersion error in white-light Linnik interferometers and its ...
    White-light interferometry is a standard optical tool with which to measure profiles of discontinuous structures such as diffractive optical elements.
  26. [26]
    [PDF] White Light Interference Microscopy System Design - HAL
    A Linnik interferometer presents the distinctive feature of possessing identical microscope objectives in the reference and object arms, in order to be able to ...
  27. [27]
    Scanning white-light interferometer. The PZT translates the object in ...
    The PZT translates the object in a direction perpendicular to the optical axis of the objective while intensity images are recorded with the detector array.
  28. [28]
    A Micro-displacement Stage for Scanning White-light Interferometry
    The stage's vertical scanning range is up to 6mm, the vertical resolution 1nm. Finite element analysis has been used to study the behavior of the flexural hinge ...<|separator|>
  29. [29]
    [PDF] Improved vertical-scanning interferometry
    We describe a method that combines phase-shifting and coherence-peak-sensing techniques to permit measurements with the height resolution of phase-shifting ...
  30. [30]
    NewView™ 9000 - ZYGO
    The NewView 9000 optical profiler features Coherence Scanning Interferometry technology for delivering sub-nanometer precision.
  31. [31]
    [PDF] Zygo NewView 200 Datasheet | ArtisanTG
    Need more Z resolution? The. NewView 200 can be upgraded to include Phase Measuring Interferometry to push the Z repeatability down to. 0.1 nm Rq!
  32. [32]
    Bruker ContourGT Optical Profiler - 3D Optical Microscope: NUANCE
    The Bruker ContourGT features white light interferometry, non-contact surface profiling, best Z resolution, largest field of view, and is a 3D optical ...Missing: specifications | Show results with:specifications
  33. [33]
    [PDF] ContourGT-K 3D Optical Microscope - INUP | IIT Madras
    The ContourGT-K offers 2D/3D measurement, high-resolution imaging, best Z resolution, large field of view, and low noise, high speed, accuracy, and precision.
  34. [34]
    [PDF] 1 HIGH-SPEED WHITE LIGHT INTERFEROMETRY FOR IMAGING ...
    The system uses white light interferometry (WLI) as a means of finding focus across the field of view. The interferometer vertically scans through different ...
  35. [35]
    High-speed lateral scanning white-light phase shift interferometry
    Jun 11, 2024 · In this study, we present lateral scanning white light interferometry (LS-WLI), where phase-shifting algorithms are applied to inspect the topography of a ...
  36. [36]
    Fundamental aspects of resolution and precision in vertical ...
    Mar 4, 2016 · We discuss the height and lateral resolution that can be achieved in vertical scanning white-light interferometry (SWLI).Missing: MEMS semiconductor
  37. [37]
    Zygo Optical Surface Profilometer
    Lateral Resolution to 0.36 microns (objective lens dependent); Sample table size: 3.5 x 8 x 8 in. Zygo Profilometer Features: Non-contact, three-dimensional, ...
  38. [38]
    InSight WLI: Automated 3D Metrology for MEMS Manufacturing - AZoM
    This allows for precise investigation of semiconductor structures and surfaces with lateral resolution as low as 200 nm. Repeatability across measurement ...
  39. [39]
  40. [40]
    Vibration effects on an environmentally tolerant scanning white light ...
    Surface metrology instruments such as the scanning white light interferometer (SWLI) normally require thermal and vibration isolation and are therefore ...Missing: temperature stabilization
  41. [41]
    Lateral scanning white-light interferometer - Optica Publishing Group
    White-light vertical scanning interferometry is a well-established technique for retrieving the three-dimensional shapes of small objects, ...
  42. [42]
    [PDF] Principles of interferometry - Eso.org
    Mar 1, 2010 · λ λ λ π λ λ π λ. The fringes are modulated with an envelope with a characteristic width equal to the coherence length, Λ coh. = λ2. 0. /Δλ. Page ...
  43. [43]
  44. [44]
    Coherence Length - RP Photonics
    The coherence length is a measure of temporal coherence, expressed as the propagation distance over which the coherence significantly decays.Missing: white | Show results with:white
  45. [45]
  46. [46]
    Advantages of Measuring Surface Roughness with White Light ...
    This application note discusses the use of mean roughness measurements with white light interferometry (WLI) optical profilers.
  47. [47]
    How White Light Interferometry Can Measure Surface Roughness
    May 14, 2020 · This article outlines the utilization of mean roughness measurements with white light interferometry (WLI) optical profilers.Missing: modes | Show results with:modes<|separator|>
  48. [48]
    3D Optical Profilers | Bruker
    White-light interferometry (WLI), also known as coherence scanning interferometry, is a non-contact optical profilometry method that takes ...
  49. [49]
    Characterizing Microlenses via White Light Interferometry - AZoM
    Jul 31, 2023 · White light interferometry (WLI) is also known as coherence scanning interferometry. ... data acquisition, such as: Autoscan ...
  50. [50]
    High-precision shape measurement by white-light interferometry ...
    White-light interferometry measures shape by detecting the coherence envelope peak. It uses real-time scanner error correction to improve accuracy by measuring ...
  51. [51]
    Characterization of micromechanical structures using white-light ...
    Instead of these methods, this paper uses white-light interferometry surface profiling for material characterization and device inspection. Interferometry ...
  52. [52]
    White Light Interferometry: A Comprehensive Guide - FindLight
    Nov 1, 2023 · Advantages of White Light Interferometry​​ WLI presents a gamut of advantages that position it as a premier technique for high-precision surface ...
  53. [53]
    3D imaging of human skin: optical in-vivo tomography and topology ...
    We discuss different modifications of white light interferometry, for the acquisition of human skin morphology. In a first experiment we display the ...
  54. [54]
    Using white light interferometry to measure etching of dental enamel
    Aug 5, 2025 · A new method has been developed where white light interferometry (WLI) is used to study the progression of dental erosions.
  55. [55]
    Surface roughness and wear performance of Bioflx versus stainless ...
    Mar 5, 2025 · Surface roughness was measured using an optical profilometer (white light interferometry, 20X objective lens). For wear measurement, seven ...
  56. [56]
    Precise measurement with white light interferometer - Micro-Epsilon
    White light interferometers use infrared light in the non-visible range (approx. 840 nm wavelength), which means that the measurement position cannot be seen.
  57. [57]
    Single-channel white light interference vibration-resistant ...
    Aug 13, 2025 · To address this issue, this project proposes a single channel white light interference anti-vibration measurement method based on non-uniform ...
  58. [58]
    Deep learning for sub-Nyquist sampling scanning white light ...
    Nov 9, 2023 · This Letter introduces sub-Nyquist sampling vertical scanning white light interferometry (SWLI) using deep learning.
  59. [59]
    High-speed lateral scanning white-light phase shift interferometry
    ### Summary of Key Advancements in Speed for White Light Interferometry
  60. [60]
    All-sapphire fiber-optic sensor for the simultaneous measurement of ...
    Apr 5, 2024 · An all-sapphire fiber-optic extrinsic Fabry-Perot interferometric (EFPI) sensor for the simultaneous measurement of ultra-high temperature ...
  61. [61]
    Simulation-Driven End-to-End Deep Learning Method for White ...
    In the field of white-light interferometry, deep neural networks have been explored for various tasks including resolution enhancement [27,28], noise reduction ...
  62. [62]
    Bruker Introduces Next-Generation 3D Areal Surface Measurement ...
    Jan 31, 2023 · Bruker Corporation today announced the release of two new white-light interferometry (WLI) systems: the ContourX-1000 and NPFLEX-1000 optical profilometers.Missing: hybrid structured
  63. [63]
    Integrated Photonics for Quantum Communications and Metrology
    ### Summary of On-Chip Implementations and Quantum-Enhanced White Light Interferometry