Interferometry

Figure 1. The light path through a Michelson interferometer. The two light rays with a common source combine at the half-silvered mirror to reach the detector. They may either interfere constructively (strengthening in intensity) if their light waves arrive in phase, or interfere destructively (weakening in intensity) if they arrive out of phase, depending on the exact distances between the three mirrors.

Interferometry is a family of techniques in which waves, usually electromagnetic, are superimposed in order to extract information about the waves.[1] Interferometry is an important investigative technique in the fields of astronomy, fiber optics, engineering metrology, optical metrology, oceanography, seismology, spectroscopy (and its applications to chemistry), quantum mechanics, nuclear and particle physics, plasma physics, remote sensing, biomolecular interactions, surface profiling, microfluidics, mechanical stress/strain measurement, and velocimetry.[2]:1–2

Interferometers are widely used in science and industry for the measurement of small displacements, refractive index changes and surface irregularities. In analytical science, interferometers are used in continuous wave Fourier transform spectroscopy to analyze light containing features of absorption or emission associated with a substance or mixture. An astronomical interferometer consists of two or more separate telescopes that combine their signals, offering a resolution equivalent to that of a telescope of diameter equal to the largest separation between its individual elements.

Basic principles

Figure 2. Formation of fringes in a Michelson interferometer
Figure 3. Colored and monochromatic fringes in a Michelson interferometer: (a) White light fringes where the two beams differ in the number of phase inversions; (b) White light fringes where the two beams have experienced the same number of phase inversions; (c) Fringe pattern using monochromatic light (sodium D lines)

Interferometry makes use of the principle of superposition to combine waves in a way that will cause the result of their combination to have some meaningful property that is diagnostic of the original state of the waves. This works because when two waves with the same frequency combine, the resulting pattern is determined by the phase difference between the two waves—waves that are in phase will undergo constructive interference while waves that are out of phase will undergo destructive interference. Most interferometers use light or some other form of electromagnetic wave.[2]:3–12

Typically (see Fig. 1, the well-known Michelson configuration) a single incoming beam of coherent light will be split into two identical beams by a beam splitter (a partially reflecting mirror). Each of these beams travels a different route, called a path, and they are recombined before arriving at a detector. The path difference, the difference in the distance traveled by each beam, creates a phase difference between them. It is this introduced phase difference that creates the interference pattern between the initially identical waves.[2]:14–17 If a single beam has been split along two paths, then the phase difference is diagnostic of anything that changes the phase along the paths. This could be a physical change in the path length itself or a change in the refractive index along the path.[2]:93–103

As seen in Fig. 2a and 2b, the observer has a direct view of mirror M1 seen through the beam splitter, and sees a reflected image M'2 of mirror M2. The fringes can be interpreted as the result of interference between light coming from the two virtual images S'1 and S'2 of the original source S. The characteristics of the interference pattern depend on the nature of the light source and the precise orientation of the mirrors and beam splitter. In Fig. 2a, the optical elements are oriented so that S'1 and S'2 are in line with the observer, and the resulting interference pattern consists of circles centered on the normal to M1 and M'2. If, as in Fig. 2b, M1 and M'2 are tilted with respect to each other, the interference fringes will generally take the shape of conic sections (hyperbolas), but if M1 and M'2 overlap, the fringes near the axis will be straight, parallel, and equally spaced. If S is an extended source rather than a point source as illustrated, the fringes of Fig. 2a must be observed with a telescope set at infinity, while the fringes of Fig. 2b will be localized on the mirrors.[2]:17

Use of white light will result in a pattern of colored fringes (see Fig. 3).[2]:26 The central fringe representing equal path length may be light or dark depending on the number of phase inversions experienced by the two beams as they traverse the optical system.[2]:26,171–172 (See Michelson interferometer for a discussion of this.)

Categories

Interferometers and interferometric techniques may be categorized by a variety of criteria:

Homodyne versus heterodyne detection

The most important and widely used application of the heterodyne technique is in the superheterodyne receiver (superhet), invented by U.S. engineer Edwin Howard Armstrong in 1918. In this circuit, the incoming radio frequency signal from the antenna is mixed with a signal from a local oscillator (LO) and converted by the heterodyne technique to a lower fixed frequency signal called the intermediate frequency (IF). This IF is amplified and filtered, before being applied to a detector which extracts the audio signal, which is sent to the loudspeaker.[4]
Optical heterodyne detection is an extension of the heterodyne technique to higher (visible) frequencies.[3]

Double path versus common path

Figure 4. Four examples of common path interferometers

Wavefront splitting versus amplitude splitting

Figure 5. Two wavefront splitting interferometers
  • In 1803, Young's interference experiment played a major role in the general acceptance of the wave theory of light. If white light is used in Young's experiment, the result is a white central band of constructive interference corresponding to equal path length from the two slits, surrounded by a symmetrical pattern of colored fringes of diminishing intensity. In addition to continuous electromagnetic radiation, Young's experiment has been performed with individual photons,[8] with electrons,[9][10] and with buckyball molecules large enough to be seen under an electron microscope.[11]
  • Lloyd's mirror generates interference fringes by combining direct light from a source (blue lines) and light from the source's reflected image (red lines) from a mirror held at grazing incidence. The result is an asymmetrical pattern of fringes. Interestingly, the band of equal path length, nearest the mirror, is dark rather than bright. In 1834, Humphrey Lloyd interpreted this effect as proof that the phase of a front-surface reflected beam is inverted.[12][13]
Figure 6. Three amplitude-splitting interferometers: Fizeau, Mach–Zehnder, and Fabry Pérot
  • The Fizeau interferometer is shown as it might be set up to test an optical flat. A precisely figured reference flat is placed on top of the flat being tested, separated by narrow spacers. The reference flat is slightly beveled (only a fraction of a degree of beveling is necessary) to prevent the rear surface of the flat from producing interference fringes. Separating the test and reference flats allows the two flats to be tilted with respect to each other. By adjusting the tilt, which adds a controlled phase gradient to the fringe pattern, one can control the spacing and direction of the fringes, so that one may obtain an easily interpreted series of nearly parallel fringes rather than a complex swirl of contour lines. Separating the plates, however, necessitates that the illuminating light be collimated. Fig 6 shows a collimated beam of monochromatic light illuminating the two flats and a beam splitter allowing the fringes to be viewed on-axis.[15][16]
  • The Mach–Zehnder interferometer is a more versatile instrument than the Michelson interferometer. Each of the well separated light paths is traversed only once, and the fringes can be adjusted so that they are localized in any desired plane.[2]:18 Typically, the fringes would be adjusted to lie in the same plane as the test object, so that fringes and test object can be photographed together. If it is decided to produce fringes in white light, then, since white light has a limited coherence length, on the order of micrometers, great care must be taken to equalize the optical paths or no fringes will be visible. As illustrated in Fig. 6, a compensating cell would be placed in the path of the reference beam to match the test cell. Note also the precise orientation of the beam splitters. The reflecting surfaces of the beam splitters would be oriented so that the test and reference beams pass through an equal amount of glass. In this orientation, the test and reference beams each experience two front-surface reflections, resulting in the same number of phase inversions. The result is that light traveling an equal optical path length in the test and reference beams produces a white light fringe of constructive interference.[17][18]
  • The heart of the Fabry–Pérot interferometer is a pair of partially silvered glass optical flats spaced several millimeters to centimeters apart with the silvered surfaces facing each other. (Alternatively, a Fabry–Pérot etalon uses a transparent plate with two parallel reflecting surfaces.)[2]:35–36 As with the Fizeau interferometer, the flats are slightly beveled. In a typical system, illumination is provided by a diffuse source set at the focal plane of a collimating lens. A focusing lens produces what would be an inverted image of the source if the paired flats were not present; i.e. in the absence of the paired flats, all light emitted from point A passing through the optical system would be focused at point A'. In Fig. 6, only one ray emitted from point A on the source is traced. As the ray passes through the paired flats, it is multiply reflected to produce multiple transmitted rays which are collected by the focusing lens and brought to point A' on the screen. The complete interference pattern takes the appearance of a set of concentric rings. The sharpness of the rings depends on the reflectivity of the flats. If the reflectivity is high, resulting in a high Q factor (i.e. high finesse), monochromatic light produces a set of narrow bright rings against a dark background.[19] In Fig. 6, the low-finesse image corresponds to a reflectivity of 0.04 (i.e. unsilvered surfaces) versus a reflectivity of 0.95 for the high-finesse image.

Applications

Physics and astronomy


Figure 7. Michelson-Morley experiment with
cryogenic optical resonators

Figure 8. Fourier transform spectroscopy

Figure 9. A picture of the solar corona taken
with the LASCO C1 coronagraph
Michelson interferometers are used in tunable narrow band optical filters[25] and as the core hardware component of Fourier transform spectrometers.[26]
When used as a tunable narrow band filter, Michelson interferometers exhibit a number of advantages and disadvantages when compared with competing technologies such as Fabry–Pérot interferometers or Lyot filters. Michelson interferometers have the largest field of view for a specified wavelength, and are relatively simple in operation, since tuning is via mechanical rotation of waveplates rather than via high voltage control of piezoelectric crystals or lithium niobate optical modulators as used in a Fabry–Pérot system. Compared with Lyot filters, which use birefringent elements, Michelson interferometers have a relatively low temperature sensitivity. On the negative side, Michelson interferometers have a relatively restricted wavelength range and require use of prefilters which restrict transmittance.[27]
Fig. 8 illustrates the operation of a Fourier transform spectrometer, which is essentially a Michelson interferometer with one mirror movable. (A practical Fourier transform spectrometer would substitute corner cube reflectors for the flat mirrors of the conventional Michelson interferometer, but for simplicity, the illustration does not show this.) An interferogram is generated by making measurements of the signal at many discrete positions of the moving mirror. A Fourier transform converts the interferogram into an actual spectrum.[28]
Fabry-Pérot thin-film etalons are used in narrow bandpass filters capable of selecting a single spectral line for imaging; for example, the H-alpha line or the Ca-K line of the Sun or stars. Fig. 10 shows an Extreme ultraviolet Imaging Telescope (EIT) image of the Sun at 195 Ångströms, corresponding to a spectral line of multiply-ionized iron atoms.[29] EIT used multilayer coated reflective mirrors that were coated with alternate layers of a light "spacer" element (such as silicon), and a heavy "scatterer" element (such as molybdenum). Approximately 100 layers of each type were placed on each mirror, with a thickness of around 10 nm each. The layer thicknesses were tightly controlled so that at the desired wavelength, reflected photons from each layer interfered constructively.
The Laser Interferometer Gravitational-Wave Observatory (LIGO) uses two 4-km Michelson-Fabry-Pérot interferometers for the detection of gravitational waves.[30] In this application, the Fabry–Pérot cavity is used to store photons for almost a millisecond while they bounce up and down between the mirrors. This increases the time a gravitational wave can interact with the light, which results in a better sensitivity at low frequencies. Smaller cavities, usually called mode cleaners, are used for spatial filtering and frequency stabilization of the main laser.
Mach-Zehnder interferometers are also used to study one of the most counterintuitive predictions of quantum mechanics, the phenomenon known as quantum entanglement.[33][34]
Figure 11. The VLA interferometer
Early radio telescope interferometers used a single baseline for measurement. Later astronomical interferometers, such as the Very Large Array illustrated in Fig 11, used arrays of telescopes arranged in a pattern on the ground. A limited number of baselines will result in insufficient coverage. This was alleviated by using the rotation of the Earth to rotate the array relative to the sky. Thus, a single baseline could measure information in multiple orientations by taking repeated measurements, a technique called Earth-rotation synthesis. Baselines thousands of kilometers long were achieved using very long baseline interferometry.[35]
Astronomical optical interferometry has had to overcome a number of technical issues not shared by radio telescope interferometry. The short wavelengths of light necessitate extreme precision and stability of construction. For example, spatial resolution of 1 milliarcsecond requires 0.5 µm stability in a 100 m baseline. Optical interferometric measurements require high sensitivity, low noise detectors that did not become available until the late 1990s. Astronomical "seeing", the turbulence that causes stars to twinkle, introduces rapid, random phase changes in the incoming light, requiring kilohertz data collection rates to be faster than the rate of turbulence.[36][37] Despite these technical difficulties, roughly a dozen astronomical optical interferometers are now in operation offering resolutions down to the fractional milliarcsecond range. Fig. 12 shows a movie assembled from aperture sythesis images of the Beta Lyrae system, a binary star system approximately 960 light-years (290 parsecs) away in the constellation Lyra, as observed by the CHARA array with the MIRC instrument. The brighter component is the primary star, or the mass donor. The fainter component is the thick disk surrounding the secondary star, or the mass gainer. The two components are separated by 1 milli-arcsecond. Tidal distortions of the mass donor and the mass gainer are both clearly visible.[38]
Electron holography is an imaging technique that photographically records the electron interference pattern of an object, which is then reconstructed to yield a greatly magnified image of the original object.[40] This technique was developed to enable greater resolution in electron microscopy than is possible using conventional imaging techniques. The resolution of conventional electron microscopy is not limited by electron wavelength, but by the large aberrations of electron lenses.[41]
Neutron interferometry has been used to investigate the Aharonov–Bohm effect, to examine the effects of gravity acting on an elementary particle, and to demonstrate a strange behavior of fermions that is at the basis of the Pauli exclusion principle: Unlike macroscopic objects, when fermions are rotated by 360° about any axis, they do not return to their original state, but develop a minus sign in their wave function. In other words, a fermion needs to be rotated 720° before returning to its original state.[42]
Atom interferometry techniques are reaching sufficient precision to allow laboratory-scale tests of general relativity.[43]

Engineering and applied science

Figure 13. Optical flat interference fringes
Figure 14. Twyman-Green Interferometer
Figure 15. Optical testing with a Fizeau interferometer and a computer generated hologram
Optical heterodyne detection is used for coherent Doppler lidar measurements capable of detecting very weak light scattered in the atmosphere and monitoring wind speeds with high accuracy. It has application in optical fiber communications, in various high resolution spectroscopic techniques, and the self-heterodyne method can be used to measure the linewidth of a laser.[3][55]
Figure 16. Frequency comb of a mode-locked laser. The dashed lines represent an extrapolation of the mode frequencies towards the frequency of the carrier–envelope offset (CEO). The vertical grey line represents an unknown optical frequency. The horizontal black lines indicate the two lowest beat frequency measurements.
Optical heterodyne detection is an essential technique used in high-accuracy measurements of the frequencies of optical sources, as well as in the stabilization of their frequencies. Until a relatively few years ago, lengthy frequency chains were needed to connect the microwave frequency of a cesium or other atomic time source to optical frequencies. At each step of the chain, a frequency multiplier would be used to produce a harmonic of the frequency of that step, which would be compared by heterodyne detection with the next step (the output of a microwave source, far infrared laser, infrared laser, or visible laser). Each measurement of a single spectral line required several years of effort in the construction of a custom frequency chain. Currently, optical frequency combs have provided a much simpler method of measuring optical frequencies. If a mode-locked laser is modulated to form a train of pulses, its spectrum is seen to consist of the carrier frequency surrounded by a closely spaced comb of optical sideband frequencies with a spacing equal to the pulse repetition frequency (Fig. 16). The pulse repetition frequency is locked to that of the frequency standard, and the frequencies of the comb elements at the red end of the spectrum are doubled and heterodyned with the frequencies of the comb elements at the blue end of the spectrum, thus allowing the comb to serve as its own reference. In this manner, locking of the frequency comb output to an atomic standard can be performed in a single step. To measure an unknown frequency, the frequency comb output is dispersed into a spectrum. The unknown frequency is overlapped with the appropriate spectral segment of the comb and the frequency of the resultant heterodyne beats is measured.[56][57]
Figure 17. Phase shifting and Coherence scanning interferometers
  • Phase Shifting Interferometry addresses several issues associated with the classical analysis of static interferograms. Classically, one measures the positions of the fringe centers. As seen in Fig. 13, fringe deviations from straightness and equal spacing provide a measure of the aberration. Errors in determining the location of the fringe centers provide the inherent limit to precision of the classical analysis, and any intensity variations across the interferogram will also introduce error. There is a trade-off between precision and number of data points: closely spaced fringes provide many data points of low precision, while widely spaced fringes provide a low number of high precision data points. Since fringe center data is all that one uses in the classical analysis, all of the other information that might theoretically be obtained by detailed analysis of the intensity variations in an interferogram is thrown away.[65][66] Finally, with static interferograms, additional information is needed to determine the polarity of the wavefront: In Fig. 13, one can see that the tested surface on the right deviates from flatness, but one cannot tell from this single image whether this deviation from flatness is concave or convex. Traditionally, this information would be obtained using non-automated means, such as by observing the direction that the fringes move when the reference surface is pushed.[67]
Phase shifting interferometry overcomes these limitations by not relying on finding fringe centers, but rather by collecting intensity data from every point of the CCD image sensor. As seen in Fig. 17, multiple interferograms (at least three) are analyzed with the reference optical surface shifted by a precise fraction of a wavelength between each exposure using a piezoelectric transducer (PZT). Alternatively, precise phase shifts can be introduced by modulating the laser frequency.[68] The captured images are processed by a computer to calculate the optical wavefront errors. The precision and reproducibility of PSI is far greater than possible in static interferogram analysis, with measurement repeatabilities of a hundredth of a wavelength being routine.[65][66] Phase shifting technology has been adapted to a variety of interferometer types such as Twyman-Green, Mach–Zehnder, laser Fizeau, and even common path configurations such as point diffraction and lateral shearing interferometers.[67][69] More generally, phase shifting techniques can be adapted to almost any system that uses fringes for measurement, such as holographic and speckle interferometry.[67]
Figure 18. Lunate cells of Nepenthes khasiana visualized by Scanning White Light Interferometry (SWLI)
Figure 19. Twyman-Green interferometer set up as a white light scanner
  • In coherence scanning interferometry,[70] interference is only achieved when the path length delays of the interferometer are matched within the coherence time of the light source. CSI monitors the fringe contrast rather than the phase of the fringes.[2]:105
Fig. 17 illustrates a CSI microscope using a Mirau interferometer in the objective; other forms of interferometer used with white light include the Michelson interferometer (for low magnification objectives, where the reference mirror in a Mirau objective would interrupt too much of the aperture) and the Linnik interferometer (for high magnification objectives with limited working distance).[71] The sample (or alternatively, the objective) is moved vertically over the full height range of the sample, and the position of maximum fringe contrast is found for each pixel.[61][72] The chief benefit of coherence scanning interferometry is that systems can be designed that do not suffer from the 2 pi ambiguity of coherent interferometry,[73][74][75] and as seen in Fig. 18, which scans a 180μm x 140μm x 10μm volume, it is well suited to profiling steps and rough surfaces. The axial resolution of the system is determined in part by the coherence length of the light source.[76][77] Industrial applications include in-process surface metrology, roughness measurement, 3D surface metrology in hard-to-reach spaces and in hostile environments, profilometry of surfaces with high aspect ratio features (grooves, channels, holes), and film thickness measurement (semi-conductor and optical industries, etc.).[78][79]
Fig. 19 illustrates a Twyman–Green interferometer set up for white light scanning of a macroscopic object.
Holographic interferometry was discovered by accident as a result of mistakes committed during the making of holograms. Early lasers were relatively weak and photographic plates were insensitive, necessitating long exposures during which vibrations or minute shifts might occur in the optical system. The resultant holograms, which showed the holographic subject covered with fringes, were considered ruined.[80]
Eventually, several independent groups of experimenters in the mid-60s realized that the fringes encoded important information about dimensional changes occurring in the subject, and began intentionally producing holographic double exposures. The main Holographic interferometry article covers the disputes over priority of discovery that occurred during the issuance of the patent for this method.[81]
Double- and multi- exposure holography is one of three methods used to create holographic interferograms. A first exposure records the object in an unstressed state. Subsequent exposures on the same photographic plate are made while the object is subjected to some stress. The composite image depicts the difference between the stressed and unstressed states.[82]
Real-time holography is a second method of creating holographic interferograms. A holograph of the unstressed object is created. This holograph is illuminated with a reference beam to generate a hologram image of the object directly superimposed over the original object itself while the object is being subjected to some stress. The object waves from this hologram image will interfere with new waves coming from the object. This technique allows real time monitoring of shape changes.[82]
The third method, time-average holography, involves creating a holograph while the object is subjected to a periodic stress or vibration. This yields a visual image of the vibration pattern.[82]
When lasers were first invented, laser speckle was considered to be a severe drawback in using lasers to illuminate objects, particularly in holographic imaging because of the grainy image produced. It was later realized that speckle patterns could carry information about the object's surface deformations. Butters and Leendertz developed the technique of speckle pattern interferometry in 1970,[85] and since then, speckle has been exploited in a variety of other applications. A photograph is made of the speckle pattern before deformation, and a second photograph is made of the speckle pattern after deformation. Digital subtraction of the two images results in a correlation fringe pattern, where the fringes represent lines of equal deformation. Short laser pulses in the nanosecond range can be used to capture very fast transient events. A phase problem exists: In the absence of other information, one cannot tell the difference between contour lines indicating a peak versus contour lines indicating a trough. To resolve the issue of phase ambiguity, ESPI may be combined with phase shifting methods.[86][87]

Biology and medicine


Figure 22. Typical optical setup of single point OCT
     
Figure 23. Central serous retinopathy,imaged using
optical coherence tomography

See also

References

Wikimedia Commons has media related to Interferometry.
  1. Bunch, Bryan H; Hellemans, Alexander (April 2004). The History of Science and Technology. Houghton Mifflin Harcourt. p. 695. ISBN 978-0-618-22123-3.
  2. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 Hariharan, P. (2007). Basics of Interferometry. Elsevier Inc. ISBN 0-12-373589-0.
  3. 1 2 3 4 Paschotta, Rüdiger. "Optical Heterodyne Detection". RP Photonics Consulting GmbH. Retrieved 1 April 2012.
  4. Poole, Ian. "The superhet or superheterodyne radio receiver". Radio-Electronics.com. Retrieved 22 June 2012.
  5. Mallick, S.; Malacara, D. (2007). "Common-Path Interferometers". Optical Shop Testing. p. 97. doi:10.1002/9780470135976.ch3. ISBN 9780470135976.
  6. Verma, R.K. (2008). Wave Optics. Discovery Publishing House. pp. 97–110. ISBN 81-8356-114-4.
  7. "Interferential Devices – Introduction". OPI – Optique pour l'Ingénieur. Retrieved 1 April 2012.
  8. Ingram Taylor, Sir Geoffrey (1909). "Interference Fringes with Feeble Light" (PDF). Proc. Cam. Phil. Soc. 15: 114. Retrieved 2 January 2013.
  9. Jönsson, C (1961). "Elektroneninterferenzen an mehreren künstlich hergestellten Feinspalten". Zeitschrift für Physik 161 (4): 454–474. Bibcode:1961ZPhy..161..454J. doi:10.1007/BF01342460.
  10. Jönsson, C (1974). "Electron diffraction at multiple slits". American Journal of Physics 4: 4–11. Bibcode:1974AmJPh..42....4J. doi:10.1119/1.1987592.
  11. Arndt, M.; Zeilinger, A. (2004). "Heisenberg's Uncertainty and Matter Wave Interferometry with Large Molecules". In Buschhorn, G. W.; Wess, J. Fundamental Physics-- Heisenberg and Beyond: Werner Heisenberg Centennial Symposium "Developments in Modern Physics". Springer. pp. 35–52. ISBN 3540202013. Retrieved 26 May 2012.
  12. Carroll, Brett. "Simple Lloyd’s Mirror" (PDF). American Association of Physics Teachers. Retrieved 5 April 2012.
  13. Serway, R.A.; Jewett, J.W. (2010). Principles of physics: a calculus-based text, Volume 1. Brooks Cole. pp. 905–905. ISBN 0-534-49143-X.
  14. Nolte, David D. (2012). Optical Interferometry for Biology and Medicine. Springer. pp. 17–26. ISBN 1-4614-0889-X.
  15. "Guideline for Use of Fizeau Interferometer in Optical Testing" (PDF). NASA. Retrieved 8 April 2012.
  16. "Interferential devices – Fizeau Interferometer". Optique pour l'Ingénieur. Retrieved 8 April 2012.
  17. Zetie, K.P.; Adams, S.F.; Tocknell, R.M. "How does a Mach–Zehnder interferometer work?" (PDF). Physics Department, Westminster School, London. Retrieved 8 April 2012.
  18. Ashkenas, Harry I. (1950). The design and construction of a Mach-Zehnder interferometer for use with the GALCIT Transonic Wind Tunnel. Engineer's thesis. California Institute of Technology.
  19. Betzler, Klaus. "Fabry-Perot Interferometer" (PDF). Fachbereich Physik, Universität Osnabrück. Retrieved 8 April 2012.
  20. Michelson, A.A.; Morley, E.W. (1887). "On the Relative Motion of the Earth and the Luminiferous Ether" (PDF). American Journal of Science 34 (203): 333–345. doi:10.2475/ajs.s3-34.203.333.
  21. Miller, Dayton C. (1933). "The Ether-Drift Experiment and the Determination of the Absolute Motion of the Earth". Reviews of Modern Physics 5 (3): 203–242. Bibcode:1933RvMP....5..203M. doi:10.1103/RevModPhys.5.203. White light fringes were chosen for the observations because they consist of a small group of fringes having a central, sharply defined black fringe which forms a permanent zero reference mark for all readings.
  22. Müller, H.; Herrmann, S.; Braxmaier, C.; Schiller, S.; Peters, A. (2003). "Modern Michelson-Morley experiment using cryogenic optical resonators". Phys. Rev. Lett. 91 (2): 020401. arXiv:physics/0305117. Bibcode:2003PhRvL..91b0401M. doi:10.1103/PhysRevLett.91.020401. PMID 12906465.
  23. Eisele, C.; Nevsky, A.; Schiller, S. (2009). "Laboratory Test of the Isotropy of Light Propagation at the 10-17 Level". Physical Review Letters 103 (9): 090401. Bibcode:2009PhRvL.103i0401E. doi:10.1103/PhysRevLett.103.090401. PMID 19792767.
  24. Herrmann, S.; Senger, A.; Möhle, K.; Nagel, M.; Kovalchuk, E.; Peters, A. (2009). "Rotating optical cavity experiment testing Lorentz invariance at the 10-17 level". Physical Review D 80 (10). arXiv:1002.1284. Bibcode:2009PhRvD..80j5011H. doi:10.1103/PhysRevD.80.105011.
  25. Scherrer, P.H.; Bogart, R.S.; Bush, R.I.; Hoeksema, J.; Kosovichev, A.G.; Schou, J. (1995). "The Solar Oscillations Investigation – Michelson Doppler Imager". Solar Physics 162: 129–188. Bibcode:1995SoPh..162..129S. doi:10.1007/BF00733429. Retrieved 2 April 2012.
  26. Stroke, G.W.; Funkhouser, A.T. (1965). "Fourier-transform spectroscopy using holographic imaging without computing and with stationary interferometers" (PDF). Physics Letters 16 (3): 272–274. Bibcode:1965PhL....16..272S. doi:10.1016/0031-9163(65)90846-2. Retrieved 2 April 2012.
  27. Gary, G.A.; Balasubramaniam, K.S. "Additional Notes Concerning the Selection of a Multiple-Etalon System for ATST" (PDF). Advanced Technology Solar Telescope. Retrieved 29 April 2012.
  28. "Spectrometry by Fourier transform". OPI – Optique pour l'Ingénieur. Retrieved 3 April 2012.
  29. "Halloween 2003 Solar Storms: SOHO/EIT Ultraviolet, 195 Ã". NASA/Goddard Space Flight Center Scientific Visualization Studio. Retrieved 20 June 2012.
  30. "LIGO -Laser Interferometer Gravitational-Wave Observatory". Caltech/MIT. Retrieved 4 April 2012.
  31. Chevalerias, R.; Latron, Y.; Veret, C. (1957). "Methods of Interferometry Applied to the Visualization of Flows in Wind Tunnels". Journal of the Optical Society of America 47 (8): 703. doi:10.1364/JOSA.47.000703.
  32. Ristić, Slavica. "Flow visualization techniques in wind tunnels – optical methods (Part II)" (PDF). Military Technical Institute, Serbia. Retrieved 6 April 2012.
  33. Paris, M.G.A. (1999). "Entanglement and visibility at the output of a Mach-Zehnder interferometer" (PDF). Physical Review A 59 (2): 1615–1621. arXiv:quant-ph/9811078. Bibcode:1999PhRvA..59.1615P. doi:10.1103/PhysRevA.59.1615. Retrieved 2 April 2012.
  34. Haack, G. R.; Förster, H.; Büttiker, M. (2010). "Parity detection and entanglement with a Mach-Zehnder interferometer". Physical Review B 82 (15). arXiv:1005.3976. Bibcode:2010PhRvB..82o5303H. doi:10.1103/PhysRevB.82.155303.
  35. 1 2 Monnier, John D (2003). "Optical interferometry in astronomy" (PDF). Reports on Progress in Physics 66 (5): 789–857. arXiv:astro-ph/0307036. Bibcode:2003RPPh...66..789M. doi:10.1088/0034-4885/66/5/203.
  36. Malbet, F.; Kern, P.; Schanen-Duport, I.; Berger, J.-P.; Rousselet-Perraut, K.; Benech, P. (1999). "Integrated optics for astronomical interferometry" (PDF). Astron. Astrophys. Suppl. Ser. 138: 1–10. arXiv:astro-ph/9907031. Bibcode:1999A&AS..138....1D. doi:10.1051/aas:1999263. Retrieved 10 April 2012.
  37. Baldwin, J.E.; Haniff, C.A. (2002). "The application of interferometry to optical astronomical imaging" (PDF). Phil. Trans. R. Soc. Lond. A 360 (1794): 969–986. Bibcode:2002RSPTA.360..969B. doi:10.1098/rsta.2001.0977. Retrieved 10 April 2012.
  38. Zhao, M.; Gies, D.; Monnier, J. D.; Thureau, N.; Pedretti, E.; Baron, F.; Merand, A.; Ten Brummelaar, T.; McAlister, H.; Ridgway, S. T.; Turner, N.; Sturmann, J.; Sturmann, L.; Farrington, C.; Goldfinger, P. J. (2008). "First Resolved Images of the Eclipsing and Interacting Binary β Lyrae". The Astrophysical Journal 684 (2): L95. arXiv:0808.0932. Bibcode:2008ApJ...684L..95Z. doi:10.1086/592146.
  39. Gerlich, S.; Eibenberger, S.; Tomandl, M.; Nimmrichter, S.; Hornberger, K.; Fagan, P. J.; Tüxen, J.; Mayor, M.; Arndt, M. (2011). "Quantum interference of large organic molecules". Nature Communications 2: 263–. doi:10.1038/ncomms1263. PMC 3104521. PMID 21468015.
  40. Lehmann, M; Lichte, H (December 2002). "Tutorial on off-axis electron holography". Microsc. Microanal. 8 (6): 447–66. Bibcode:2002MiMic...8..447L. doi:10.1017/S1431927602029938. PMID 12533207.
  41. Tonomura, A. (1999). Electron Holography, 2nd ed. Springer. ISBN 3-540-64555-1.
  42. Klein, T. (2009). "Neutron interferometry: A tale of three continents". Europhysics News 40 (6): 24–20. doi:10.1051/epn/2009802.
  43. Dimopoulos, S.; Graham, P.W.; Hogan, J.M.; Kasevich, M.A. (2008). "General Relativistic Effects in Atom Interferometry" (PDF). Phys. Rev. D 78 (042003). arXiv:0802.4098. Bibcode:2008PhRvD..78d2003D. doi:10.1103/PhysRevD.78.042003.
  44. Mariani, Z.; Strong, K.; Wolff, M.; et al. (2012). "Infrared measurements in the Arctic using two Atmospheric Emitted Radiance Interferometers". Atmos. Meas. Tech. 5: 329–344. Bibcode:2012AMT.....5..329M. doi:10.5194/amt-5-329-2012.
  45. Mantravadi, M. V.; Malacara, D. (2007). "Newton, Fizeau, and Haidinger Interferometers". Optical Shop Testing. p. 1. doi:10.1002/9780470135976.ch1. ISBN 9780470135976.
  46. Malacara, D. (2007). "Twyman–Green Interferometer". Optical Shop Testing. p. 46. doi:10.1002/9780470135976.ch2. ISBN 9780470135976.
  47. Michelson, A. A. (1918). "On the Correction of Optical Surfaces". Proceedings of the National Academy of Sciences of the United States of America 4 (7): 210–212. doi:10.1073/pnas.4.7.210. PMC 1091444. PMID 16576300.
  48. "Interferential Devices – Twyman-Green Interferometer". OPI – Optique pour l'Ingénieur. Retrieved 4 April 2012.
  49. Heideman, R. G.; Kooyman, R. P. H.; Greve, J. (1993). "Performance of a highly sensitive optical waveguide Mach-Zehnder interferometer immunosensor". Sensors and Actuators B: Chemical 10 (3): 209–217. doi:10.1016/0925-4005(93)87008-D.
  50. Oliver, W. D.; Yu, Y.; Lee, J. C.; Berggren, K. K.; Levitov, L. S.; Orlando, T. P. (2005). "Mach-Zehnder Interferometry in a Strongly Driven Superconducting Qubit". Science 310 (5754): 1653–1657. doi:10.1126/science.1119678. PMID 16282527.
  51. Nieradko, Ł.; Gorecki, C.; JóZwik, M.; Sabac, A.; Hoffmann, R.; Bertz, A. (2006). "Fabrication and optical packaging of an integrated Mach-Zehnder interferometer on top of a movable micromirror". Journal of Microlithography, Microfabrication, and Microsystems 5 (2): 023009. doi:10.1117/1.2203366.
  52. Burge, J. H.; Zhao, C.; Dubin, M. (2010). "Measurement of aspheric mirror segments using Fizeau interferometry with CGH correction" (PDF). Proceedings of SPIE 7739: 773902. doi:10.1117/12.857816.
  53. Anderson, R.; Bilger, H.R.; Stedman, G.E. (1994). ""Sagnac effect" A century of Earth-rotated interferometers" (PDF). Am. J. Phys. 62 (11): 975–985. Bibcode:1994AmJPh..62..975A. doi:10.1119/1.17656. Retrieved 30 March 2012.
  54. Golio, Mike (2007). RF and Microwave Applications and Systems. CRC Press. pp. 14.1–14.17. ISBN 0849372194. Retrieved 27 June 2012.
  55. Paschotta, Rüdiger. "Self-heterodyne Linewidth Measurement". RP Photonics. Retrieved 22 June 2012.
  56. "Optical Frequency Comb". National Research Council, Canada. Retrieved 23 June 2012.
  57. Paschotta, Rüdiger. "Frequency Combs". RP Photonics. Retrieved 23 June 2012.
  58. Schmit, J. (1993). "Spatial and temporal phase-measurement techniques: a comparison of major error sources in one dimension". Proceedings of SPIE 1755. pp. 202–201. doi:10.1117/12.140770.
  59. Larkin, K.G. (1996). "Efficient nonlinear algorithm for envelope detection in white light interferometry" (PDF). Journal of the Optical Society of America 13 (4): 832–843. Bibcode:1996JOSAA..13..832L. doi:10.1364/JOSAA.13.000832. Retrieved 1 April 2012.
  60. ISO. (2013). 25178-604:2013(E): Geometrical product specification (GPS) – Surface texture: Areal – Nominal characteristics of non-contact (coherence scanning interferometric microscopy) instruments (2013(E) ed.). Geneva: International Organization for Standardization.
  61. 1 2 Harasaki, A.; Schmit, J.; Wyant, J. C. (2000). "Improved vertical-scanning interferometry" (PDF). Applied Optics 39 (13): 2107–2115. Bibcode:2000ApOpt..39.2107H. doi:10.1364/AO.39.002107. Retrieved 21 May 2012.
  62. De Groot, P (2015). "Principles of interference microscopy for the measurement of surface topography". Advances in Optics and Photonics 7: 1–65. doi:10.1364/AOP.7.000001.
  63. 1 2 3 Olszak, A.G.; Schmit, J.; Heaton, M.G. "Interferometry: Technology and Applications" (PDF). Bruker. Retrieved 1 April 2012.
  64. de Groot, Peter; Deck, Leslie (1995). "Surface Profiling by Analysis of White-light Interferograms in the Spatial Frequency Domain". Journal of Modern Optics 42 (2): 389–401. doi:10.1080/09500349514550341.
  65. 1 2 "Phase-Shifting Interferometry for Determining Optical Surface Quality". Newport Corporation. Retrieved 12 May 2012.
  66. 1 2 "How Phase Interferometers work". Graham Optical Systems. 2011. Retrieved 12 May 2012.
  67. 1 2 3 Schreiber, H.; Bruning, J. H. (2007). "Phase Shifting Interferometry". Optical Shop Testing. p. 547. doi:10.1002/9780470135976.ch14. ISBN 9780470135976.
  68. Sommargren, G. E. (1986). US Patent 4,594,003.
  69. Ferraro, P.; Paturzo, M.; Grilli, S. (2007). "Optical wavefront measurement using a novel phase-shifting point-diffraction interferometer". SPIE. Retrieved 26 May 2012.
  70. P. de Groot, J., "Interference Microscopy for Surface Structure Analysis," in Handbook of Optical Metrology, edited by T. Yoshizawa, chapt.31, pp. 791-828, (CRC Press, 2015).
  71. Schmit, J.; Creath, K.; Wyant, J. C. (2007). "Surface Profilers, Multiple Wavelength, and White Light Intereferometry". Optical Shop Testing. p. 667. doi:10.1002/9780470135976.ch15. ISBN 9780470135976.
  72. "HDVSI – Introducing High Definition Vertical Scanning Interferometry for Nanotechnology Research from Veeco Instruments". Veeco. Retrieved 21 May 2012.
  73. Plucinski, J.; Hypszer, R.; Wierzba, P.; Strakowski, M.; Jedrzejewska-Szczerska, M.; Maciejewski, M.; Kosmowski, B.B. (2008). "Optical low-coherence interferometry for selected technical applications" (PDF). Bulletin of the Polish Academy of Sciences 56 (2): 155–172. Retrieved 8 April 2012.
  74. Yang, C.-H.; Wax, A; Dasari, R.R.; Feld, M.S. (2002). "2π ambiguity-free optical distance measurement with subnanometer precision with a novel phase-crossing low-coherence interferometer" (PDF). Optics Letters 27 (2): 77–79. Bibcode:2002OptL...27...77Y. doi:10.1364/OL.27.000077.
  75. Hitzenberger, C. K.; Sticker, M.; Leitgeb, R.; Fercher, A. F. (2001). "Differential phase measurements in low-coherence interferometry without 2pi ambiguity". Optics Letters 26 (23): 1864–1866. doi:10.1364/ol.26.001864. PMID 18059719.
  76. Wojtek J. Walecki, Kevin Lai, Vitalij Souchkov, Phuc Van, SH Lau, Ann Koo physica status solidi (c) Volume 2, Issue 3, Pages 984–989
  77. W. J. Walecki et al. "Non-contact fast wafer metrology for ultra-thin patterned wafers mounted on grinding and dicing tapes" Electronics Manufacturing Technology Symposium, 2004. IEEE/CPMT/SEMI 29th International Volume, Issue, July 14–16, 2004 Page(s): 323–325
  78. "Coating Thickness Measurement". Lumetrics, Inc. Retrieved 28 October 2013.
  79. "Typical profilometry measurements". Novacam Technologies, Inc. Retrieved 25 June 2012.
  80. "Holographic interferometry". Oquagen. 2008. Retrieved 22 May 2012.
  81. Hecht, Jeff (1998). Laser, Light of a Million Uses. Dover Publications, Inc. pp. 229–230. ISBN 0-486-40193-6.
  82. 1 2 3 Fein, H (September 1997). "Holographic Interferometry: Nondestructive tool" (PDF). The Industrial Physicist: 37–39.
  83. "PIA01762: Space Radar Image of Kilauea, Hawaii". NASA/JPL. 1999. Retrieved 17 June 2012.
  84. Jones R & Wykes C, Holographic and Speckle Interferometry, 1989, Cambridge University Press
  85. Butters, J. N.; Leendertz, J. A. (1971). "A double exposure technique for speckle pattern interferometry". Journal of Physics E: Scientific Instruments 4 (4): 277–279. doi:10.1088/0022-3735/4/4/004.
  86. Dvořáková, P.; Bajgar, V.; Trnka, J. (2007). "Dynamic Electronic Speckle Pattern Interferometry in Application to Measure Out-Of-Plane Displacement" (PDF). Engineering Mechanics 14 (1/2): 37–44.
  87. Moustafa, N. A.; Hendawi, N. (2003). "Comparative Phase-Shifting Digital Speckle Pattern Interferometry Using Single Reference Beam Technique" (PDF). Egypt. J. Sol. 26 (2): 225–229. Retrieved 22 May 2012.
  88. Buga, A.; Jokela, J.; Putrimas, R. "Traceability, stability and use of the Kyviskes calibration baseline–the first 10 years". Environmental Engineering, The 7th International Conference (PDF). Vilnius Gediminas Technical University. pp. 1274–1280. Retrieved 9 April 2012.
  89. Huang, D.; Swanson, E.A.; Lin, C.P.; Schuman, J.S.; Stinson, W.G.; Chang, W.; Hee, M.R.; Flotte, T.; Gregory, K.; Puliafito, C.A.; Fujimoto, J.G. (1991). "Optical Coherence Tomography" (PDF). Science 254 (5035): 1178–81. Bibcode:1991Sci...254.1178H. doi:10.1126/science.1957169. PMID 1957169. Retrieved 10 April 2012.
  90. Fercher, A.F. (1996). "Optical Coherence Tomography" (PDF). Journal of Biomedical Optics 1 (2): 157–173. Bibcode:1996JBO.....1..157F. doi:10.1117/12.231361. Retrieved 10 April 2012.
  91. Lang, Walter. "Nomarski Differential Interference-Contrast Microscopy" (PDF). Carl Zeiss, Oberkochen. Retrieved 10 April 2012.
  92. Wax, A.; Pyhtila, J. W.; Graf, R. N.; Nines, R.; Boone, C. W.; Dasari, R. R.; Feld, M. S.; Steele, V. E.; Stoner, G. D. (2005). "Prospective grading of neoplastic change in rat esophagus epithelium using angle-resolved low-coherence interferometry". Journal of Biomedical Optics 10 (5): 051604. doi:10.1117/1.2102767. PMID 16292952.
  93. Pyhtila, J. W.; Chalut, K. J.; Boyer, J. D.; Keener, J.; d'Amico, T.; Gottfried, M.; Gress, F.; Wax, A. (2007). "In situ detection of nuclear atypia in Barrett's esophagus by using angle-resolved low-coherence interferometry". Gastrointestinal Endoscopy 65 (3): 487–491. doi:10.1016/j.gie.2006.10.016. PMID 17321252.
  94. Fitzgerald, Richard (2000). "Phase-sensitive x-ray imaging". Physics Today 53 (7): 23. Bibcode:2000PhT....53g..23F. doi:10.1063/1.1292471.
  95. 1 2 David, C, Nohammer, B, Solak, H H, & Ziegler E (2002). "Differential x-ray phase contrast imaging using a shearing interferometer". Applied Physics Letters 81 (17): 3287–3289. Bibcode:2002ApPhL..81.3287D. doi:10.1063/1.1516611.
  96. Wilkins, S W, Gureyev, T E, Gao, D, Pogany, A & Stevenson, A W (1996). "Phase-contrast imaging using polychromatic hard X-rays". Nature 384 (6607): 335–338. Bibcode:1996Natur.384..335W. doi:10.1038/384335a0.
  97. Davis, T J, Gao, D, Gureyev, T E, Stevenson, A W & Wilkins, S W (1995). "Phase-contrast imaging of weakly absorbing materials using hard X-rays". Nature 373 (6515): 595–598. Bibcode:1995Natur.373..595D. doi:10.1038/373595a0.
  98. Momose, A, Takeda, T, Itai, Y & Hirano, K (1996). "Phase-contrast X-ray computed tomography for observing biological soft tissues". Nature Medicine 2 (4): 473–475. doi:10.1038/nm0496-473. PMID 8597962.
This article is issued from Wikipedia - version of the Wednesday, February 10, 2016. The text is available under the Creative Commons Attribution/Share Alike but additional terms may apply for the media files.