Speckle (interference)

Last updated

Speckle, speckle pattern, or speckle noise designates the granular structure observed in coherent light, resulting from random interference. Speckle patterns are used in a wide range of metrology techniques, as they generally allow high sensitivity and simple setups. They can also be a limiting factor in imaging systems, such as radar, synthetic aperture radar (SAR), medical ultrasound and optical coherence tomography. [1] [2] [3] [4] Speckle is not external noise; rather, it is an inherent fluctuation in diffuse reflections, because the scatterers are not identical for each cell, and the coherent illumination wave is highly sensitive to small variations in phase changes. [5]

Contents

Speckle patterns arise when coherent light is randomised. The simplest case of such randomisation is when light reflects off an optically rough surface. Optically rough means that the surface profile contains fluctuations larger than the wavelength. Most common surfaces are rough to visible light, such as paper, wood, or paint.

The vast majority of surfaces, synthetic or natural, are extremely rough on the scale of the wavelength. We see the origin of this phenomenon if we model our reflectivity function as an array of scatterers. Because of the finite resolution, at any time we are receiving from a distribution of scatterers within the resolution cell. These scattered signals add coherently; that is, they add constructively and destructively depending on the relative phases of each scattered waveform. Speckle results from these patterns of constructive and destructive interference shown as bright and dark dots in the image. [6]

Speckle in conventional radar increases the mean grey level of a local area. [7] Speckle in SAR is generally serious, causing difficulties for image interpretation. [7] [8] It is caused by coherent processing of backscattered signals from multiple distributed targets. In SAR oceanography, for example, speckle is caused by signals from elementary scatterers, the gravity-capillary ripples, and manifests as a pedestal image, beneath the image of the sea waves. [9] [10]

The speckle can also represent some useful information, particularly when it is linked to the laser speckle and to the dynamic speckle phenomenon, where the changes of the spatial speckle pattern over time can be used as a measurement of the surface's activity, such as which is useful for measuring displacement fields via digital image correlation.


History

Although scientists have investigated this phenomenon since the time of Newton, [11] speckles have come into prominence since the invention of the laser. [12]

Formation

The speckle effect is a result of the interference of many waves of the same frequency, having different phases and amplitudes, which add together to give a resultant wave whose amplitude, and therefore intensity, varies randomly. If we model each wave by a vector, we can then see that if we add a number of vectors with random angles together, the length of the resulting vector can be anything from zero to the sum of the individual vector lengths—a 2-dimensional random walk, sometimes known as a drunkard's walk. In the limit of many interfering waves, and for polarised waves, the distribution of intensities (which go as the square of the vector's length) becomes exponential , where is the mean intensity. [1] [2] [13] [14]

When a surface is illuminated by a light wave, according to diffraction theory, each point on an illuminated surface acts as a source of secondary spherical waves. The light at any point in the scattered light field is made up of waves which have been scattered from each point on the illuminated surface. If the surface is rough enough to create path-length differences exceeding one wavelength, giving rise to phase changes greater than 2π, the amplitude, and hence the intensity, of the resultant light varies randomly.

If light of low coherence (i.e., made up of many wavelengths) is used, a speckle pattern will not normally be observed, because the speckle patterns produced by individual wavelengths have different dimensions and will normally average one another out. However, we can observe speckle patterns in polychromatic light in some conditions, such as with partially coherent light, [15] or spatially coherent light. [16]

Types

Subjective speckles

Laser speckle on a digital camera image from a green laser pointer. This is a subjective speckle pattern. (Note that the color differences in the image are introduced by limitations of the camera system.) Laser speckle.jpg
Laser speckle on a digital camera image from a green laser pointer. This is a subjective speckle pattern. (Note that the color differences in the image are introduced by limitations of the camera system.)

When a rough surface which is illuminated by a coherent light (e.g. a laser beam) is imaged, a speckle pattern is observed in the image plane; this is called a "subjective speckle pattern" – see image above. It is called "subjective" because the detailed structure of the speckle pattern depends on the viewing system parameters; for instance, if the size of the lens aperture changes, the size of the speckles change. If the position of the imaging system is altered, the pattern will gradually change and will eventually be unrelated to the original speckle pattern.

We can explain this as follows. We can consider each point in the image to be illuminated by a finite area in the object.[ clarification needed ] We determine the size of this area by the diffraction-limited resolution of the lens which is given by the Airy disk whose diameter is 2.4λu/D, where λ is the wavelength of the light, u is the distance between the object and the lens, and D is the diameter of the lens aperture. (This is a simplified model of diffraction-limited imaging.)

The light at neighboring points in the image has been scattered from areas which have many points in common and the intensity of two such points will not differ much. However, two points in the image which are illuminated by areas in the object which are separated by the diameter of the Airy disk, have light intensities which are unrelated. This corresponds to a distance in the image of 2.4λv/D where v is the distance between the lens and the image. Thus, the "size" of the speckles in the image is of this order.

We can observe the change in speckle size with lens aperture by looking at a laser spot on a wall directly, and then through a very small hole. The speckles will be seen to increase significantly in size. Also, the speckle pattern itself will change when moving the position of the eye while keeping the laser pointer steady. A further proof that the speckle pattern is formed only in the image plane (in the specific case the eye's retina) is that the speckles will stay visible if the eye's focus is shifted away from the wall (this is different for an objective speckle pattern, where the speckle visibility is lost under defocusing).

Objective speckles

A photograph of an objective speckle pattern. This is the light field formed when a laser beam was scattered from a plastic surface onto a wall. Objective speckle.jpg
A photograph of an objective speckle pattern. This is the light field formed when a laser beam was scattered from a plastic surface onto a wall.

When laser light which has been scattered off a rough surface falls on another surface, it forms an "objective speckle pattern". If a photographic plate or another 2-D optical sensor is located within the scattered light field without a lens, a speckle pattern is obtained whose characteristics depend on the geometry of the system and the wavelength of the laser. The speckle pattern in the figure was obtained by pointing a laser beam at the surface of a mobile phone so that the scattered light fell onto an adjacent wall. A photograph was then taken of the speckle pattern formed on the wall. Strictly speaking, this also has a second subjective speckle pattern but its dimensions are much smaller than the objective pattern so it is not seen in the image.

Contributions from the whole of the scattering surface make up the light at a given point in the speckle pattern. The relative phases of these scattered waves vary across the scattering surface, so that the resulting phase on each point of the second surface varies randomly. The pattern is the same regardless of how it is imaged, just as if it were a painted pattern.

The "size" of the speckles is a function of the wavelength of the light, the size of the laser beam which illuminates the first surface, and the distance between this surface and the surface where the speckle pattern is formed. This is the case because when the angle of scattering changes such that the relative path difference between light scattered from the centre of the illuminated area compared with light scattered from the edge of the illuminated area changes by λ, the intensity becomes uncorrelated. Dainty [1] derives an expression for the mean speckle size as λz/L where L is the width of the illuminated area and z is the distance between the object and the location of the speckle pattern.

Near-field speckles

Objective speckles are usually obtained in the far field (also called Fraunhofer region, that is the zone where Fraunhofer diffraction happens). This means that they are generated "far" from the object that emits or scatters light. We can also observe speckles close to the scattering object, in the near field (also called Fresnel region, that is, the region where Fresnel diffraction happens). We call this kind of speckles near-field speckles. See near and far field for a more rigorous definition of "near" and "far".

The statistical properties of a far-field speckle pattern (i.e., the speckle form and dimension) depend on the form and dimension of the region hit by laser light. By contrast, a very interesting feature of near field speckles is that their statistical properties are closely related to the form and structure of the scattering object: objects that scatter at high angles generate small near field speckles, and vice versa. Under Rayleigh–Gans condition, in particular, speckle dimension mirrors the average dimension of the scattering objects, while, in general, the statistical properties of near field speckles generated by a sample depend on the light scattering distribution. [17] [18]

Actually, the condition under which the near field speckles appear has been described as more strict than the usual Fresnel condition. [19]

Applications


Speckle patterns have been used in a variety of applications in microscopy, [20] [21] imaging, [22] [23] and optical manipulation. [24] [25] [26]

When lasers were first invented, the speckle effect was considered to be a severe drawback in using lasers to illuminate objects, particularly in holographic imaging because of the grainy image produced. Researchers later realized that speckle patterns could carry information about the object's surface deformations, and exploited this effect in holographic interferometry and electronic speckle pattern interferometry. [27] Speckle imaging and eye testing using speckle also use the speckle effect.

Speckle is the chief limitation of coherent lidar and coherent imaging in optical heterodyne detection.

In the case of near field speckles, the statistical properties depend on the light scattering distribution of a given sample. This allows the use of near field speckle analysis to detect the scattering distribution; this is the so-called near-field scattering technique. [28]

When the speckle pattern changes in time, due to changes in the illuminated surface, the phenomenon is known as dynamic speckle, and it can be used to measure activity, by means of, for example, an optical flow sensor (optical computer mouse). In biological materials, the phenomenon is known as biospeckle.

In a static environment, changes in speckle can also be used as a sensitive probe of the light source. This can be used in a wavemeter configuration, with a resolution around 1 attometre, [29] (equivalent to 1 part in 1012 of the wavelength, equivalent to measuring the length of a football field at the resolution of a single atom [30] ) and can also stabilise the wavelength of lasers [31] or measure polarization. [32]

The disordered pattern produced by speckle has been used in quantum simulations with cold atoms. The randomly-distributed regions of bright and dark light act as an analog of disorder in solid-state systems, and are used to investigate localization phenomena. [33]

In fluorescence microscopy, a sub-diffraction-limited resolution can be obtained in 2D from saturable/photoconvertible pattern illumination techniques like stimulated emission depletion (STED) microscopy, ground state depletion (GSD) microscopy, and reversible saturable optical fluorescence transitions (RESOLFT). Adapting speckle patterns for use in these applications enables parallel 3D super-resolution imaging. [34]

Mitigation

A green laser pointer. Reduction of the speckle was necessary to photograph the laser's Gaussian profile, accomplished by removing all lenses and projecting it onto an opaque liquid (milk) being the only surface flat and smooth enough. Green laser pointer TEM00 profile.JPG
A green laser pointer. Reduction of the speckle was necessary to photograph the laser's Gaussian profile, accomplished by removing all lenses and projecting it onto an opaque liquid (milk) being the only surface flat and smooth enough.

Speckle is considered to be a problem in laser based display systems like the Laser TV. Speckle is usually quantified by the speckle contrast. Speckle contrast reduction is essentially the creation of many independent speckle patterns, so that they average out on the retina/detector. This can be achieved by, [35]

Rotating diffusers—which destroys the spatial coherence of the laser light—can also be used to reduce the speckle. Moving/vibrating screens or fibers may also be solutions. [36] The Mitsubishi Laser TV appears to use such a screen which requires special care according to their product manual. A more detailed discussion on laser speckle reduction can be found here. [37]

Synthetic array heterodyne detection was developed to reduce speckle noise in coherent optical imaging and coherent differential absorption LIDAR.

Signal processing methods

In scientific applications, a spatial filter can be used to reduce speckle.

Several different methods are used to eliminate speckle, based upon different mathematical models of the phenomenon. [9] One method, for example, employs multiple-look processing (a.k.a. multi-look processing), averaging out the speckle by taking several "looks" at a target in a single radar sweep. [7] [8] The average is the incoherent average of the looks. [8]

A second method involves using adaptive and non-adaptive filters on the signal processing (where adaptive filters adapt their weightings across the image to the speckle level, and non-adaptive filters apply the same weightings uniformly across the entire image). Such filtering also eliminates actual image information as well, in particular high-frequency information, and the applicability of filtering and the choice of filter type involves tradeoffs. Adaptive speckle filtering is better at preserving edges and detail in high-texture areas (such as forests or urban areas). Non-adaptive filtering is simpler to implement, and requires less computational power, however. [7] [8]

There are two forms of non-adaptive speckle filtering: one based on the mean and one based upon the median (within a given rectangular area of pixels in the image). The latter is better at preserving edges whilst eliminating spikes, than the former is. There are many forms of adaptive speckle filtering, [38] including the Lee filter, the Frost filter, and the refined gamma maximum-A-posteriori (RGMAP) filter. They all rely upon three fundamental assumptions in their mathematical models, however: [7]

The Lee filter converts the multiplicative model into an additive one, thereby reducing the problem of dealing with speckle to a known tractable case. [39]

Wavelet analysis

Recently, the use of wavelet transform has led to significant advances in image analysis. The main reason for the use of multiscale processing is the fact that many natural signals, when decomposed into wavelet bases are significantly simplified and can be modeled by known distributions. Besides, wavelet decomposition is able to separate signals at different scales and orientations. Therefore, the original signal at any scale and direction can be recovered and useful details are not lost. [40]

The first multiscale speckle reduction methods were based on the thresholding of detail subband coefficients. [41] Wavelet thresholding methods have some drawbacks: (i) the choice of threshold is made in an ad hoc manner, supposing that wanted and unwanted components of the signal obey their known distributions, irrespective of their scale and orientations; and (ii) the thresholding procedure generally results in some artifacts in the denoised image. To address these disadvantages, non-linear estimators, based on Bayes' theory were developed. [40] [42]

Analogies

Speckle patterns can also be observed over time instead of space. This is the case of phase sensitive optical time-domain reflectometry, where multiple reflections of a coherent pulse generated at different instants interfere to produce a pseudorandom time-domain signal. [43]

Optical vortices in speckle patterns

Speckle interference pattern may be decomposed in the sum of plane waves. There exist a set of points where amplitude of electromagnetic field is exactly zero. Researchers had recognized these points as dislocations of wave trains. [44] We know these phase dislocations of electromagnetic fields as optical vortices.

There is a circular energy flow around each vortex core. Thus each vortex in the speckle pattern carries optical angular momentum. The angular momentum density is given by: [45]

Typically vortices appear in speckle pattern in pairs. These vortex - antivortex pairs are placed randomly in space. One may show that electromagnetic angular momentum of each vortex pair is close to zero. [46] In phase conjugating mirrors based on stimulated Brillouin scattering optical vortices excite acoustical vortices. [47]

Apart from formal decomposition in Fourier series the speckle pattern may be composed for plane waves emitted by tilted regions of the phase plate. This approach significantly simplifies numerical modelling. 3D numerical emulation demonstrates the intertwining of vortices which leads to formation of ropes in optical speckle. [48]

See also

Related Research Articles

<span class="mw-page-title-main">Diffraction</span> Phenomenon of the motion of waves

Diffraction is the interference or bending of waves around the corners of an obstacle or through an aperture into the region of geometrical shadow of the obstacle/aperture. The diffracting object or aperture effectively becomes a secondary source of the propagating wave. Italian scientist Francesco Maria Grimaldi coined the word diffraction and was the first to record accurate observations of the phenomenon in 1660.

<span class="mw-page-title-main">Polarization (waves)</span> Property of waves that can oscillate with more than one orientation

Polarization is a property of transverse waves which specifies the geometrical orientation of the oscillations. In a transverse wave, the direction of the oscillation is perpendicular to the direction of motion of the wave. A simple example of a polarized transverse wave is vibrations traveling along a taut string (see image); for example, in a musical instrument like a guitar string. Depending on how the string is plucked, the vibrations can be in a vertical direction, horizontal direction, or at any angle perpendicular to the string. In contrast, in longitudinal waves, such as sound waves in a liquid or gas, the displacement of the particles in the oscillation is always in the direction of propagation, so these waves do not exhibit polarization. Transverse waves that exhibit polarization include electromagnetic waves such as light and radio waves, gravitational waves, and transverse sound waves in solids.

<span class="mw-page-title-main">Raman spectroscopy</span> Spectroscopic technique

Raman spectroscopy is a spectroscopic technique typically used to determine vibrational modes of molecules, although rotational and other low-frequency modes of systems may also be observed. Raman spectroscopy is commonly used in chemistry to provide a structural fingerprint by which molecules can be identified.

<span class="mw-page-title-main">Interferometry</span> Measurement method using interference of waves

Interferometry is a technique which uses the interference of superimposed waves to extract information. Interferometry typically uses electromagnetic waves and is an important investigative technique in the fields of astronomy, fiber optics, engineering metrology, optical metrology, oceanography, seismology, spectroscopy, quantum mechanics, nuclear and particle physics, plasma physics, biomolecular interactions, surface profiling, microfluidics, mechanical stress/strain measurement, velocimetry, optometry, and making holograms.

<span class="mw-page-title-main">Optical tweezers</span> Scientific instruments

Optical tweezers are scientific instruments that use a highly focused laser beam to hold and move microscopic and sub-microscopic objects like atoms, nanoparticles and droplets, in a manner similar to tweezers. If the object is held in air or vacuum without additional support, it can be called optical levitation.

<span class="mw-page-title-main">Michelson interferometer</span> Common configuration for optical interferometry

The Michelson interferometer is a common configuration for optical interferometry and was invented by the 19/20th-century American physicist Albert Abraham Michelson. Using a beam splitter, a light source is split into two arms. Each of those light beams is reflected back toward the beamsplitter which then combines their amplitudes using the superposition principle. The resulting interference pattern that is not directed back toward the source is typically directed to some type of photoelectric detector or camera. For different applications of the interferometer, the two light paths can be with different lengths or incorporate optical elements or even materials under test.

Particle image velocimetry (PIV) is an optical method of flow visualization used in education and research. It is used to obtain instantaneous velocity measurements and related properties in fluids. The fluid is seeded with tracer particles which, for sufficiently small particles, are assumed to faithfully follow the flow dynamics. The fluid with entrained particles is illuminated so that particles are visible. The motion of the seeding particles is used to calculate speed and direction of the flow being studied.

<span class="mw-page-title-main">Coherent backscattering</span> Scattering of a radiation wave as it travels through a complex medium

In physics, coherent backscattering is observed when coherent radiation propagates through a medium which has a large number of scattering centers of size comparable to the wavelength of the radiation.

Holographic interferometry (HI) is a technique which enables the measurements of static and dynamic displacements of objects with optically rough surfaces at optical interferometric precision. These measurements can be applied to stress, strain and vibration analysis, as well as to non-destructive testing and radiation dosimetry. It can also be used to detect optical path length variations in transparent media, which enables, for example, fluid flow to be visualised and analyzed. It can also be used to generate contours representing the form of the surface.

<span class="mw-page-title-main">Coherent diffraction imaging</span> Lensless computational imaging method

Coherent diffractive imaging (CDI) is a "lensless" technique for 2D or 3D reconstruction of the image of nanoscale structures such as nanotubes, nanocrystals, porous nanocrystalline layers, defects, potentially proteins, and more. In CDI, a highly coherent beam of X-rays, electrons or other wavelike particle or photon is incident on an object.

<span class="mw-page-title-main">Electronic speckle pattern interferometry</span>

Electronic speckle pattern interferometry (ESPI), also known as TV holography, is a technique that uses laser light, together with video detection, recording and processing, to visualise static and dynamic displacements of components with optically rough surfaces. The visualisation is in the form of fringes on the image, where each fringe normally represents a displacement of half a wavelength of the light used.

Optical heterodyne detection is a method of extracting information encoded as modulation of the phase, frequency or both of electromagnetic radiation in the wavelength band of visible or infrared light. The light signal is compared with standard or reference light from a "local oscillator" (LO) that would have a fixed offset in frequency and phase from the signal if the latter carried null information. "Heterodyne" signifies more than one frequency, in contrast to the single frequency employed in homodyne detection.

<span class="mw-page-title-main">Ptychography</span>

Ptychography is a computational method of microscopic imaging. It generates images by processing many coherent interference patterns that have been scattered from an object of interest. Its defining characteristic is translational invariance, which means that the interference patterns are generated by one constant function moving laterally by a known amount with respect to another constant function. The interference patterns occur some distance away from these two components, so that the scattered waves spread out and "fold" into one another as shown in the figure.

Ultrasound-modulated optical tomography (UOT), also known as Acousto-Optic Tomography (AOT), is a hybrid imaging modality that combines light and sound; it is a form of tomography involving ultrasound. It is used in imaging of biological soft tissues and has potential applications for early cancer detection. As a hybrid modality which uses both light and sound, UOT provides some of the best features of both: the use of light provides strong contrast and sensitivity ; these two features are derived from the optical component of UOT. The use of ultrasound allows for high resolution, as well as a high imaging depth. However, the difficulty of tackling the two fundamental problems with UOT have caused UOT to evolve relatively slowly; most work in the field is limited to theoretical simulations or phantom / sample studies.

A random laser (RL) is a laser in which optical feedback is provided by scattering particles. As in conventional lasers, a gain medium is required for optical amplification. However, in contrast to Fabry–Pérot cavities and distributed feedback lasers, neither reflective surfaces nor distributed periodic structures are used in RLs, as light is confined in an active region by diffusive elements that either may or may not be spatially distributed inside the gain medium.

The technique of vibrational analysis with scanning probe microscopy allows probing vibrational properties of materials at the submicrometer scale, and even of individual molecules. This is accomplished by integrating scanning probe microscopy (SPM) and vibrational spectroscopy. This combination allows for much higher spatial resolution than can be achieved with conventional Raman/FTIR instrumentation. The technique is also nondestructive, requires non-extensive sample preparation, and provides more contrast such as intensity contrast, polarization contrast and wavelength contrast, as well as providing specific chemical information and topography images simultaneously.

<span class="mw-page-title-main">Orbital angular momentum of light</span> Type of angular momentum in light

The orbital angular momentum of light (OAM) is the component of angular momentum of a light beam that is dependent on the field spatial distribution, and not on the polarization. OAM can be split into two types. The internal OAM is an origin-independent angular momentum of a light beam that can be associated with a helical or twisted wavefront. The external OAM is the origin-dependent angular momentum that can be obtained as cross product of the light beam position and its total linear momentum.

<span class="mw-page-title-main">Electromagnetic metasurface</span>

An electromagnetic metasurface refers to a kind of artificial sheet material with sub-wavelength thickness. Metasurfaces can be either structured or unstructured with subwavelength-scaled patterns in the horizontal dimensions.

Optical holography is a technique which enables an optical wavefront to be recorded and later re-constructed. Holography is best known as a method of generating three-dimensional images but it also has a wide range of other applications.

Laser speckle contrast imaging (LSCI), also called laser speckle imaging (LSI), is an imaging modality based on the analysis of the blurring effect of the speckle pattern. The operation of LSCI is having a wide-field illumination of a rough surface through a coherent light source. Then using photodetectors such as CCD camera or CMOS sensors imaging the resulting laser speckle pattern caused by the interference of coherent light. In biomedical use, the coherent light is typically in the red or near-infrared region to ensure higher penetration depth. When scattering particles moving during the time, the interference caused by the coherent light will have fluctuations which will lead to the intensity variations detected via the photodetector, and this change of the intensity contain the information of scattering particles' motion. Through image the speckle patterns with finite exposure time, areas with scattering particles will appear blurred.

References

  1. 1 2 3 Dainty, C., ed. (1984). Laser Speckle and Related Phenomena (2nd ed.). Springer-Verlag. ISBN   978-0-387-13169-6.
  2. 1 2 Goodman, J. W. (1976). "Some fundamental properties of speckle". JOSA. 66 (11): 1145–1150. Bibcode:1976JOSA...66.1145G. doi:10.1364/josa.66.001145.
  3. Hua, Tao; Xie, Huimin; Wang, Simon; Hu, Zhenxing; Chen, Pengwan; Zhang, Qingming (2011). "Evaluation of the quality of a speckle pattern in the digital image correlation method by mean subset fluctuation". Optics & Laser Technology. 43 (1): 9–13. Bibcode:2011OptLT..43....9H. doi:10.1016/j.optlastec.2010.04.010.
  4. Lecompte, D.; Smits, A.; Bossuyt, Sven; Sol, H.; Vantomme, J.; Hemelrijck, D. Van; Habraken, A.M. (2006). "Quality assessment of speckle patterns for digital image correlation". Optics and Lasers in Engineering. 44 (11): 1132–1145. Bibcode:2006OptLE..44.1132L. doi:10.1016/j.optlaseng.2005.10.004. hdl: 2268/15779 .
  5. Moreira, Alberto; Prats-Iraola, Pau; Younis, Marwan; Krieger, Gerhard; Hajnsek, Irena; Papathanassiou, Konstantinos P. (2013). "A Tutorial on Synthetic Aperture Radar" (PDF). IEEE Geoscience and Remote Sensing Magazine. 1: 6–43. doi:10.1109/MGRS.2013.2248301. S2CID   7487291.
  6. M. Forouzanfar and H. Abrishami-Moghaddam, Ultrasound Speckle Reduction in the Complex Wavelet Domain, in Principles of Waveform Diversity and Design, M. Wicks, E. Mokole, S. Blunt, R. Schneible, and V. Amuso (eds.), SciTech Publishing, 2010, Section B - Part V: Remote Sensing, pp. 558-77.
  7. 1 2 3 4 5 6 7 8 Brandt Tso & Paul Mather (2009). Classification Methods for Remotely Sensed Data (2nd ed.). CRC Press. pp. 37–38. ISBN   9781420090727.
  8. 1 2 3 4 Giorgio Franceschetti & Riccardo Lanari (1999). Synthetic aperture radar processing. Electronic engineering systems series. CRC Press. pp. 145 et seq. ISBN   9780849378997.
  9. 1 2 Mikhail B. Kanevsky (2008). Radar imaging of the ocean waves. Elsevier. p. 138. ISBN   9780444532091.
  10. Alexander Ya Pasmurov & Julius S. Zinoviev (2005). Radar imaging and holography. IEE radar, sonar and navigation series. Vol. 19. IET. p. 175. ISBN   9780863415029.
  11. Françon, Maurice (2012). Laser speckle and applications in optics. Elsevier.
  12. Rigden, J. D.; Gordon, E. I. (1962). "The granularity of scattered optical maser light". Proc. IRE.
  13. Bender, Nicholas; Yılmaz, Hasan; Bromberg, Yaron; Cao, Hui (2019-11-01). "Creating and controlling complex light". APL Photonics. 4 (11): 110806. arXiv: 1906.11698 . Bibcode:2019APLP....4k0806B. doi: 10.1063/1.5132960 .
  14. Bender, Nicholas; Yılmaz, Hasan; Bromberg, Yaron; Cao, Hui (2018-05-20). "Customizing speckle intensity statistics". Optica. 5 (5): 595–600. arXiv: 1711.11128 . Bibcode:2018Optic...5..595B. doi:10.1364/OPTICA.5.000595. ISSN   2334-2536. S2CID   119357011.
  15. McKechnie, T.S. (1976). "Image-plane speckle in partially coherent illumination". Optical and Quantum Electronics. 8: 61–67. doi:10.1007/bf00620441. S2CID   122771512.
  16. Facchin, M. (2023). On speckle patterns: integrating spheres, metrology, and beyond (PhD thesis). The University of St Andrews.
  17. Giglio, M.; Carpineti, M.; Vailati, A. (2000). "Space Intensity Correlations in the Near Field of the Scattered Light: A Direct Measurement of the Density Correlation Function g(r)". Physical Review Letters. 85 (7): 1416–1419. Bibcode:2000PhRvL..85.1416G. doi:10.1103/PhysRevLett.85.1416. PMID   10970518. S2CID   19689982.
  18. Giglio, M.; Carpineti, M.; Vailati, A.; Brogioli, D. (2001). "Near-Field Intensity Correlations of Scattered Light". Applied Optics. 40 (24): 4036–40. Bibcode:2001ApOpt..40.4036G. doi:10.1364/AO.40.004036. PMID   18360438.
  19. Cerbino, R. (2007). "Correlations of light in the deep Fresnel region: An extended Van Cittert and Zernike theorem" (PDF). Physical Review A. 75 (5): 053815. Bibcode:2007PhRvA..75e3815C. doi:10.1103/PhysRevA.75.053815.
  20. Ventalon, Cathie; Mertz, Jerome (2006-08-07). "Dynamic speckle illumination microscopy with translated versus randomized speckle patterns". Optics Express. 14 (16): 7198–7309. Bibcode:2006OExpr..14.7198V. doi: 10.1364/oe.14.007198 . ISSN   1094-4087. PMID   19529088.
  21. Pascucci, M.; Ganesan, S.; Tripathi, A.; Katz, O.; Emiliani, V.; Guillon, M. (2019-03-22). "Compressive three-dimensional super-resolution microscopy with speckle-saturated fluorescence excitation". Nature Communications. 10 (1): 1327. Bibcode:2019NatCo..10.1327P. doi: 10.1038/s41467-019-09297-5 . ISSN   2041-1723. PMC   6430798 . PMID   30902978.
  22. Katz, Ori; Bromberg, Yaron; Silberberg, Yaron (2009-09-28). "Compressive ghost imaging". Applied Physics Letters. 95 (13): 131110. arXiv: 0905.0321 . Bibcode:2009ApPhL..95m1110K. doi:10.1063/1.3238296. ISSN   0003-6951. S2CID   118516184.
  23. Dunn, Andrew K.; Bolay, Hayrunnisa; Moskowitz, Michael A.; Boas, David A. (2001-03-01). "Dynamic Imaging of Cerebral Blood Flow Using Laser Speckle". Journal of Cerebral Blood Flow & Metabolism. 21 (3): 195–201. doi: 10.1097/00004647-200103000-00002 . ISSN   0271-678X. PMID   11295873.
  24. Bechinger, Clemens; Di Leonardo, Roberto; Löwen, Hartmut; Reichhardt, Charles; Volpe, Giorgio; Volpe, Giovanni (2016-11-23). "Active Particles in Complex and Crowded Environments". Reviews of Modern Physics. 88 (4): 045006. arXiv: 1602.00081 . Bibcode:2016RvMP...88d5006B. doi:10.1103/revmodphys.88.045006. hdl: 11693/36533 . ISSN   0034-6861. S2CID   14940249.
  25. Volpe, Giorgio; Volpe, Giovanni; Gigan, Sylvain (2014-02-05). "Brownian Motion in a Speckle Light Field: Tunable Anomalous Diffusion and Selective Optical Manipulation". Scientific Reports. 4 (1): 3936. arXiv: 1304.1433 . Bibcode:2014NatSR...4E3936V. doi:10.1038/srep03936. ISSN   2045-2322. PMC   3913929 . PMID   24496461.
  26. Volpe, Giorgio; Kurz, Lisa; Callegari, Agnese; Volpe, Giovanni; Gigan, Sylvain (2014-07-28). "Speckle optical tweezers: micromanipulation with random light fields". Optics Express. 22 (15): 18159–18167. arXiv: 1403.0364 . Bibcode:2014OExpr..2218159V. doi:10.1364/OE.22.018159. hdl: 11693/12625 . ISSN   1094-4087. PMID   25089434. S2CID   14121619.
  27. Jones & Wykes, Robert & Catherine (1989). Holographic and Speckle Interferometry. Cambridge University Press. ISBN   9780511622465.
  28. Brogioli, D.; Vailati, A.; Giglio, M. (2002). "Heterodyne near-field scattering". Applied Physics Letters. 81 (22): 4109–11. arXiv: physics/0305102 . Bibcode:2002ApPhL..81.4109B. doi:10.1063/1.1524702. S2CID   119087994.
  29. Bruce, Graham D.; O’Donnell, Laura; Chen, Mingzhou; Dholakia, Kishan (2019-03-15). "Overcoming the speckle correlation limit to achieve a fiber wavemeter with attometer resolution". Optics Letters. 44 (6): 1367–1370. arXiv: 1909.00666 . Bibcode:2019OptL...44.1367B. doi:10.1364/OL.44.001367. ISSN   0146-9592. PMID   30874652. S2CID   78095181.
  30. Tudhope, Christine (7 March 2019). "New research could revolutionise fiber-optic communications". Phys.org. Retrieved 2019-03-08.
  31. Metzger, Nikolaus Klaus; Spesyvtsev, Roman; Bruce, Graham D.; Miller, Bill; Maker, Gareth T.; Malcolm, Graeme; Mazilu, Michael; Dholakia, Kishan (2017-06-05). "Harnessing speckle for a sub-femtometre resolved broadband wavemeter and laser stabilization". Nature Communications. 8: 15610. arXiv: 1706.02378 . Bibcode:2017NatCo...815610M. doi:10.1038/ncomms15610. PMC   5465361 . PMID   28580938.
  32. Facchin, Morgan; Bruce, Graham D.; Dholakia, Kishan; Dholakia, Kishan; Dholakia, Kishan (2020-05-15). "Speckle-based determination of the polarisation state of single and multiple laser beams". OSA Continuum. 3 (5): 1302–1313. arXiv: 2003.14408 . doi: 10.1364/OSAC.394117 . ISSN   2578-7519.
  33. Billy, Juliette; Josse, Vincent; Zuo, Zhanchun; Bernard, Alain; Hambrecht, Ben; Lugan, Pierre; Clément, David; Sanchez-Palencia, Laurent; Bouyer, Philippe (2008-06-12). "Direct observation of Anderson localization of matter waves in a controlled disorder". Nature. 453 (7197): 891–894. arXiv: 0804.1621 . Bibcode:2008Natur.453..891B. doi:10.1038/nature07000. ISSN   0028-0836. PMID   18548065. S2CID   4427739.
  34. Bender, Nicholas; Sun, Mengyuan; Yılmaz, Hasan; Bewersdorf, Joerg; Bewersdorf, Joerg; Cao, Hui (2021-02-20). "Circumventing the optical diffraction limit with customized speckles". Optica. 8 (2): 122–129. arXiv: 2007.15491 . Bibcode:2021Optic...8..122B. doi: 10.1364/OPTICA.411007 . ISSN   2334-2536.
  35. Trisnadi, Jahja I. (2002). "Speckle contrast reduction in laser projection displays". In Wu, Ming H (ed.). Projection Displays VIII. Vol. 4657. pp. 131–137. doi:10.1117/12.463781. S2CID   30764926.
  36. "Despeckler". Fiberguide. Retrieved 24 May 2019.
  37. Chellappan, Kishore V.; Erden, Erdem; Urey, Hakan (2010). "Laser-based displays: A review". Applied Optics. 49 (25): F79–98. Bibcode:2010ApOpt..49F..79C. doi:10.1364/ao.49.000f79. PMID   20820205. S2CID   3073667.
  38. Argenti, F.; Lapini, A.; Bianchi, T.; Alparone, L. (September 2013). "A Tutorial on Speckle Reduction in Synthetic Aperture Radar Images" (PDF). IEEE Geoscience and Remote Sensing Magazine. 1 (3): 6–35. doi:10.1109/MGRS.2013.2277512. S2CID   38021146.
  39. Piero Zamperoni (1995). "Image Enhancement". In Peter W. Hawkes; Benjamin Kazan; Tom Mulvey (eds.). Advances in imaging and electron physics. Vol. 92. Academic Press. p. 13. ISBN   9780120147342.
  40. 1 2 M. Forouzanfar, H. Abrishami-Moghaddam, and M. Gity, "A new multiscale Bayesian algorithm for speckle reduction in medical ultrasound images," Signal, Image and Video Processing, Springer, vol. 4, pp. 359-75, Sep. 2010
  41. Mallat, S.: A Wavelet Tour of Signal Processing. Academic Press, London (1998)
  42. Argenti, F.; Bianchi, T.; Lapini, A.; Alparone, L. (January 2012). "Fast MAP Despeckling Based on Laplacian–Gaussian Modeling of Wavelet Coefficients". IEEE Geoscience and Remote Sensing Letters. 9 (1): 13–17. Bibcode:2012IGRSL...9...13A. doi:10.1109/LGRS.2011.2158798. S2CID   25396128.
  43. Garcia-Ruiz, Andres (2016). "Speckle Analysis Method for Distributed Detection of Temperature Gradients With Φ OTDR". IEEE Photonics Technology Letters. 28 (18): 2000. Bibcode:2016IPTL...28.2000G. doi:10.1109/LPT.2016.2578043. S2CID   25243784.
  44. Nye, J. F.; Berry, M. V. (1974). "Dislocations in Wave Trains". Proceedings of the Royal Society A. 336 (1605): 165–190. Bibcode:1974RSPSA.336..165N. doi:10.1098/rspa.1974.0012. S2CID   122947659.
  45. Optical Angular Momentum
  46. Okulov, A. Yu. (2008). "Optical and sound helical structures in a Mandelstam-Brillouin mirror". JETP Letters. 88 (8): 487–491. Bibcode:2008JETPL..88..487O. doi:10.1134/S0021364008200046. S2CID   120371573.
  47. Okulov, A Yu (2008). "Angular momentum of photons and phase conjugation". Journal of Physics B. 41 (10): 101001. arXiv: 0801.2675 . Bibcode:2008JPhB...41j1001O. doi:10.1088/0953-4075/41/10/101001. S2CID   13307937.
  48. Okulov, A. Yu (2009). "Twisted speckle entities inside wave-front reversal mirrors". Physical Review A. 80 (1): 013837. arXiv: 0903.0057 . Bibcode:2009PhRvA..80a3837O. doi:10.1103/PhysRevA.80.013837. S2CID   119279889.

Further reading