Diffraction: Diffraction Refers To Various Phenomena That Occur When A Wave

Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Diffraction

Diffraction refers to various phenomena that occur when a wave


encounters an obstacle or opening. It is defined as the 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.[1][2]

In classical physics, the diffraction phenomenon is described by the


Huygens–Fresnel principle that treats each point in a propagating
wavefront as a collection of individual spherical wavelets.[3] The
characteristic bending pattern is most pronounced when a wave A diffraction pattern of a red laser
from a coherent source (such as a laser) encounters a slit/aperture beam projected onto a plate after
that is comparable in size to its wavelength, as shown in the passing through a small circular
inserted image. This is due to the addition, or interference, of aperture in another plate
different points on the wavefront (or, equivalently, each wavelet)
that travel by paths of different lengths to the registering
surface. If there are multiple, closely spaced openings (e.g., a
diffraction grating), a complex pattern of varying intensity
can result.

These effects also occur when a light wave travels through a


medium with a varying refractive index, or when a sound
wave travels through a medium with varying acoustic
impedance – all waves diffract, including gravitational
waves, water waves, and other electromagnetic waves such
as X-rays and radio waves. Furthermore, quantum mechanics
also demonstrates that matter possesses wave-like properties,
and hence, undergoes diffraction (which is measurable at
subatomic to molecular levels).[4] Infinitely many points (three shown) along
length d project phase contributions from
the wavefront, producing a continuously
varying intensity θ on the registering plate.
Contents
History
Mechanism
Examples
Single-slit diffraction
Diffraction grating
Circular aperture
General aperture
Propagation of a laser beam
Diffraction-limited imaging
Speckle patterns
Babinet's principle
"Knife edge"
Patterns
Particle diffraction
Bragg diffraction
Coherence
Applications
Diffraction before destruction
See also
References
External links

History
The effects of diffraction of light were first carefully observed and
characterized by Francesco Maria Grimaldi, who also coined the
term diffraction, from the Latin diffringere, 'to break into pieces',
referring to light breaking up into different directions. The results of
Grimaldi's observations were published posthumously in
1665.[5][6][7] Isaac Newton studied these effects and attributed Thomas Young's sketch of two-slit
them to inflexion of light rays. James Gregory (1638–1675) diffraction for water waves, which he
observed the diffraction patterns caused by a bird feather, which presented to the Royal Society in
was effectively the first diffraction grating to be discovered.[8] 1803.
Thomas Young performed a celebrated experiment in 1803
demonstrating interference from two closely spaced slits.[9]
Explaining his results by interference of the waves emanating from the two different slits, he deduced that
light must propagate as waves. Augustin-Jean Fresnel did more definitive studies and calculations of
diffraction, made public in 1816[10] and 1818,[11] and thereby gave great support to the wave theory of
light that had been advanced by Christiaan Huygens[12] and reinvigorated by Young, against Newton's
particle theory.

Mechanism
In classical physics diffraction arises because of the way in which
waves propagate; this is described by the Huygens–Fresnel
principle and the principle of superposition of waves. The
propagation of a wave can be visualized by considering every
particle of the transmitted medium on a wavefront as a point source
for a secondary spherical wave. The wave displacement at any
subsequent point is the sum of these secondary waves. When
waves are added together, their sum is determined by the relative
Photograph of single-slit diffraction in phases as well as the amplitudes of the individual waves so that the
a circular ripple tank summed amplitude of the waves can have any value between zero
and the sum of the individual amplitudes. Hence, diffraction
patterns usually have a series of maxima and minima.
In the modern quantum mechanical understanding of light propagation through a slit (or slits) every photon
has what is known as a wavefunction. The wavefunction is determined by the physical surroundings such
as slit geometry, screen distance and initial conditions when the photon is created. In important experiments
(A low-intensity double-slit experiment was first performed by G. I. Taylor in 1909, see double-slit
experiment) the existence of the photon's wavefunction was demonstrated. In the quantum approach the
diffraction pattern is created by the probability distribution, the observation of light and dark bands is the
presence or absence of photons in these areas, where these particles were more or less likely to be detected.
The quantum approach has some striking similarities to the Huygens-Fresnel principle; based on that
principle, as light travels through slits and boundaries, secondary, point light sources are created near or
along these obstacles, and the resulting diffraction pattern is going to be the intensity profile based on the
collective interference of all these lights sources that have different optical paths. That is similar to
considering the limited regions around the slits and boundaries where photons are more likely to originate
from, in the quantum formalism, and calculating the probability distribution. This distribution is directly
proportional to the intensity, in the classical formalism.

There are various analytical models which allow the diffracted field to be calculated, including the
Kirchhoff-Fresnel diffraction equation which is derived from the wave equation,[13] the Fraunhofer
diffraction approximation of the Kirchhoff equation which applies to the far field and the Fresnel diffraction
approximation which applies to the near field. Most configurations cannot be solved analytically, but can
yield numerical solutions through finite element and boundary element methods.

It is possible to obtain a qualitative understanding of many diffraction phenomena by considering how the
relative phases of the individual secondary wave sources vary, and in particular, the conditions in which the
phase difference equals half a cycle in which case waves will cancel one another out.

The simplest descriptions of diffraction are those in which the situation can be reduced to a two-
dimensional problem. For water waves, this is already the case; water waves propagate only on the surface
of the water. For light, we can often neglect one direction if the diffracting object extends in that direction
over a distance far greater than the wavelength. In the case of light shining through small circular holes we
will have to take into account the full three-dimensional nature of the problem.


Computer Generation of an Computational Optical diffraction


generated interference pattern model of an pattern ( laser),
intensity pattern from two-slit interference (analogous to X-ray
formed on a diffraction. pattern from crystallography)
screen by two-slit
diffraction from diffraction.
a square
aperture.

Colors seen in a spider


web are partially due to
diffraction, according to
some analyses.[14]

Examples
The effects of diffraction are often seen in everyday life. The most
striking examples of diffraction are those that involve light; for
example, the closely spaced tracks on a CD or DVD act as a
diffraction grating to form the familiar rainbow pattern seen when
looking at a disc. This principle can be extended to engineer a
grating with a structure such that it will produce any diffraction
pattern desired; the hologram on a credit card is an example.
Diffraction in the atmosphere by small particles can cause a bright
ring to be visible around a bright light source like the sun or the
moon. A shadow of a solid object, using light from a compact Circular waves generated by
source, shows small fringes near its edges. The speckle pattern diffraction from the narrow entrance
of a flooded coastal quarry
which is observed when laser light falls on an optically rough
surface is also a diffraction phenomenon. When deli meat appears
to be iridescent, that is diffraction off the meat fibers.[15] All these
effects are a consequence of the fact that light propagates as a wave.

Diffraction can occur with any kind of wave. Ocean waves diffract around jetties and other obstacles.
Sound waves can diffract around objects, which is why one can still hear someone calling even when
hiding behind a tree.[16]
Diffraction can also be a concern in some technical applications; it sets a
fundamental limit to the resolution of a camera, telescope, or
microscope.

Other examples of diffraction are considered below.

Single-slit diffraction

A long slit of infinitesimal width which is illuminated by light


diffracts the light into a series of circular waves and the wavefront
which emerges from the slit is a cylindrical wave of uniform
intensity, in accordance with Huygens–Fresnel principle. A solar glory on steam from hot
springs. A glory is an optical
A slit that is wider than a wavelength produces interference effects phenomenon produced by light
in the space downstream of the slit. These can be explained by backscattered (a combination of
assuming that the slit behaves as though it has a large number of diffraction, reflection and refraction)
point sources spaced evenly across the width of the slit. The towards its source by a cloud of
analysis of this system is simplified if we consider light of a single uniformly sized water droplets.
wavelength. If the incident light is coherent, these sources all have
the same phase. Light incident at a given point in the space
downstream of the slit is made up of contributions from each of
these point sources and if the relative phases of these contributions
vary by 2π or more, we may expect to find minima and maxima in
the diffracted light. Such phase differences are caused by
differences in the path lengths over which contributing rays reach
the point from the slit.

We can find the angle at which a first minimum is obtained in the


diffracted light by the following reasoning. The light from a source
located at the top edge of the slit interferes destructively with a
source located at the middle of the slit, when the path difference
between them is equal to λ/2. Similarly, the source just below the 2D Single-slit diffraction with width
top of the slit will interfere destructively with the source located just changing animation
below the middle of the slit at the same angle. We can continue this
reasoning along the entire height of the slit to conclude that the
condition for destructive interference for the entire slit is the same
as the condition for destructive interference between two narrow
slits a distance apart that is half the width of the slit. The path

difference is approximately so that the minimum intensity


occurs at an angle θmin given by

where Numerical approximation of


diffraction pattern from a slit of width
d is the width of the slit, four wavelengths with an incident
is the angle of incidence at which the minimum plane wave. The main central beam,
intensity occurs, and nulls, and phase reversals are
is the wavelength of the light apparent.
A similar argument can be used to show that if we imagine the slit
to be divided into four, six, eight parts, etc., minima are obtained at
angles θn given by

where
Graph and image of single-slit
diffraction.
n is an integer other than zero.

There is no such simple argument to enable us to find the maxima


of the diffraction pattern. The intensity profile can be calculated using the Fraunhofer diffraction equation
as

where

is the intensity at a given angle,


is the intensity at the central maximum ( ), which is also a normalization factor of the
intensity profile that can be determined by an integration from to and
conservation of energy.

is the unnormalized sinc function.

This analysis applies only to the far field (Fraunhofer diffraction), that is, at a distance much larger than the
width of the slit.

From the intensity profile above, if , the intensity will have little dependency on , hence the
wavefront emerging from the slit would resemble a cylindrical wave with azimuthal symmetry; If ,
only would have appreciable intensity, hence the wavefront emerging from the slit would resemble
that of geometrical optics.

When the incident angle of the light onto the slit is non-zero (which causes a change in the path length),
the intensity profile in the Fraunhofer regime (i.e. far field) becomes:

The choice of plus/minus sign depends on the definition of the incident angle .

Diffraction grating

A diffraction grating is an optical component with a regular pattern.


The form of the light diffracted by a grating depends on the
structure of the elements and the number of elements present, but
all gratings have intensity maxima at angles θm which are given by 2-slit (top) and 5-slit diffraction of red
the grating equation laser light
A diffraction pattern of a 633 nm
laser through a grid of 150 slits
Diffraction of a red laser using a
diffraction grating.

where

θi is the angle at which the light is incident,


d is the separation of grating elements, and
m is an integer which can be positive or negative.

The light diffracted by a grating is found by summing the light diffracted from each of the elements, and is
essentially a convolution of diffraction and interference patterns.

The figure shows the light diffracted by 2-element and 5-element gratings where the grating spacings are
the same; it can be seen that the maxima are in the same position, but the detailed structures of the
intensities are different.

Circular aperture

The far-field diffraction of a plane wave incident on a circular


aperture is often referred to as the Airy Disk. The variation in
intensity with angle is given by

,
A computer-generated image of an
where a is the radius of the circular aperture, k is equal to 2π/λ and Airy disk.
J1 is a Bessel function. The smaller the aperture, the larger the spot
size at a given distance, and the greater the divergence of the
diffracted beams.

General aperture

The wave that emerges from a point source has amplitude at location r that is given by the solution of the
frequency domain wave equation for a point source (the Helmholtz equation),

where is the 3-dimensional delta function. The delta function has only radial dependence, so the
Laplace operator (a.k.a. scalar Laplacian) in the spherical coordinate system simplifies to (see del in
cylindrical and spherical coordinates)
By direct substitution, the solution to this equation can be readily
shown to be the scalar Green's function, which in the spherical
coordinate system (and using the physics time convention ) is:

Computer generated light diffraction


This solution assumes that the delta function source is located at the pattern from a circular aperture of
origin. If the source is located at an arbitrary source point, denoted diameter 0.5 micrometre at a
by the vector and the field point is located at the point , then we wavelength of 0.6 micrometre (red-
may represent the scalar Green's function (for arbitrary source light) at distances of 0.1 cm – 1 cm
location) as: in steps of 0.1 cm. One can see the
image moving from the Fresnel
region into the Fraunhofer region
where the Airy pattern is seen.

Therefore, if an electric field, Einc(x,y) is incident on the aperture,


the field produced by this aperture distribution is given by the surface integral:

where the source point in the aperture is given by the


vector

In the far field, wherein the parallel rays


approximation can be employed, the Green's
function,

On the calculation of Fraunhofer region fields

simplifies to

as can be seen in the figure to the right (click to enlarge).

The expression for the far-zone (Fraunhofer region) field becomes

Now, since
and

the expression for the Fraunhofer region field from a planar aperture now becomes,

Letting,

and

the Fraunhofer region field of the planar aperture assumes the form of a Fourier transform

In the far-field / Fraunhofer region, this becomes the spatial Fourier transform of the aperture distribution.
Huygens' principle when applied to an aperture simply says that the far-field diffraction pattern is the spatial
Fourier transform of the aperture shape, and this is a direct by-product of using the parallel-rays
approximation, which is identical to doing a plane wave decomposition of the aperture plane fields (see
Fourier optics).

Propagation of a laser beam

The way in which the beam profile of a laser beam changes as it propagates is determined by diffraction.
When the entire emitted beam has a planar, spatially coherent wave front, it approximates Gaussian beam
profile and has the lowest divergence for a given diameter. The smaller the output beam, the quicker it
diverges. It is possible to reduce the divergence of a laser beam by first expanding it with one convex lens,
and then collimating it with a second convex lens whose focal point is coincident with that of the first lens.
The resulting beam has a larger diameter, and hence a lower divergence. Divergence of a laser beam may
be reduced below the diffraction of a Gaussian beam or even reversed to convergence if the refractive
index of the propagation media increases with the light intensity.[17] This may result in a self-focusing
effect.

When the wave front of the emitted beam has perturbations, only the transverse coherence length (where
the wave front perturbation is less than 1/4 of the wavelength) should be considered as a Gaussian beam
diameter when determining the divergence of the laser beam. If the transverse coherence length in the
vertical direction is higher than in horizontal, the laser beam divergence will be lower in the vertical
direction than in the horizontal.

Diffraction-limited imaging
The ability of an imaging system to resolve detail is ultimately
limited by diffraction. This is because a plane wave incident on a
circular lens or mirror is diffracted as described above. The light is
not focused to a point but forms an Airy disk having a central spot
in the focal plane whose radius (as measured to the first null) is

where λ is the wavelength of the light and N is the f-number (focal


length f divided by aperture diameter D) of the imaging optics; this
is strictly accurate for N≫1 (paraxial case). In object space, the
The Airy disk around each of the
corresponding angular resolution is
stars from the 2.56 m telescope
aperture can be seen in this lucky
image of the binary star zeta Boötis.

where D is the diameter of the entrance pupil of the imaging lens


(e.g., of a telescope's main mirror).

Two point sources will each produce an Airy pattern – see the photo of a binary star. As the point sources
move closer together, the patterns will start to overlap, and ultimately they will merge to form a single
pattern, in which case the two point sources cannot be resolved in the image. The Rayleigh criterion
specifies that two point sources are considered "resolved" if the separation of the two images is at least the
radius of the Airy disk, i.e. if the first minimum of one coincides with the maximum of the other.

Thus, the larger the aperture of the lens compared to the wavelength, the finer the resolution of an imaging
system. This is one reason astronomical telescopes require large objectives, and why microscope objectives
require a large numerical aperture (large aperture diameter compared to working distance) in order to obtain
the highest possible resolution.

Speckle patterns

The speckle pattern seen when using a laser pointer is another diffraction phenomenon. It is a result of the
superposition of many waves with different phases, which are produced when a laser beam illuminates a
rough surface. They add together to give a resultant wave whose amplitude, and therefore intensity, varies
randomly.

Babinet's principle

Babinet's principle is a useful theorem stating that the diffraction pattern from an opaque body is identical to
that from a hole of the same size and shape, but with differing intensities. This means that the interference
conditions of a single obstruction would be the same as that of a single slit.

"Knife edge"

The knife-edge effect or knife-edge diffraction is a truncation of a portion of the incident radiation that
strikes a sharp well-defined obstacle, such as a mountain range or the wall of a building.
The knife-edge
effect is explained by Huygens–Fresnel principle, which states that a well-defined obstruction to an
electromagnetic wave acts as a secondary source, and creates a new wavefront. This new wavefront
propagates into the geometric shadow area of the obstacle.
Knife-edge diffraction is an outgrowth of the "half-plane problem", originally solved by Arnold
Sommerfeld using a plane wave spectrum formulation. A generalization of the half-plane problem is the
"wedge problem", solvable as a boundary value problem in cylindrical coordinates. The solution in
cylindrical coordinates was then extended to the optical regime by Joseph B. Keller, who introduced the
notion of diffraction coefficients through his geometrical theory of diffraction (GTD). Pathak and
Kouyoumjian extended the (singular) Keller coefficients via the uniform theory of diffraction (UTD).

Diffraction on a sharp metallic edge Diffraction on a soft aperture, with a


gradient of conductivity over the image
width

Patterns
Several qualitative observations can be made of diffraction in general:

The angular spacing of the features in the diffraction pattern is inversely proportional to the
dimensions of the object causing the diffraction. In other words: The smaller the diffracting
object, the 'wider' the resulting diffraction pattern, and vice versa. (More precisely, this is true
of the sines of the angles.)
The diffraction angles are invariant under scaling; that is, they depend only on the ratio of the
wavelength to the size of the diffracting object.
When the diffracting object has a periodic structure, for example in a diffraction grating, the
features generally become sharper. The third figure, for example, shows a comparison of a
double-slit pattern with a pattern formed by five slits, both sets of slits having the same
spacing, between the center of one slit and the next.

Particle diffraction
According to quantum theory every particle exhibits wave properties. In particular, massive particles can
interfere with themselves and therefore diffract. Diffraction of electrons and neutrons stood as one of the
powerful arguments in favor of quantum mechanics. The wavelength associated with a particle is the de
Broglie wavelength
where h is Planck's constant and p is the momentum of the particle
(mass × velocity for slow-moving particles).

For most macroscopic objects, this wavelength is so short that it is


not meaningful to assign a wavelength to them. A sodium atom
traveling at about 30,000 m/s would have a De Broglie wavelength
of about 50 pico meters.

Because the wavelength for even the smallest of macroscopic


objects is extremely small, diffraction of matter waves is only
visible for small particles, like electrons, neutrons, atoms and small
molecules. The short wavelength of these matter waves makes
them ideally suited to study the atomic crystal structure of solids
and large molecules like proteins.

Relatively larger molecules like buckyballs were also shown to


diffract.[18]

Bragg diffraction The upper half of this image shows a


diffraction pattern of He-Ne laser
beam on an elliptic aperture. The
Diffraction from a three-dimensional periodic structure such as
lower half is its 2D Fourier transform
atoms in a crystal is called Bragg diffraction.
It is similar to what
approximately reconstructing the
occurs when waves are scattered from a diffraction grating. Bragg shape of the aperture.
diffraction is a consequence of interference between waves
reflecting from different crystal planes.
The condition of
constructive interference is given by Bragg's law:

where

λ is the wavelength,
d is the distance between crystal planes,
θ is the angle of the diffracted wave.
and m is an integer known as the order of the diffracted
beam.

Bragg diffraction may be carried out using either electromagnetic Following Bragg's law, each dot (or
radiation of very short wavelength like X-rays or matter waves like reflection) in this diffraction pattern
neutrons (and electrons) whose wavelength is on the order of (or forms from the constructive
much smaller than) the atomic spacing.[19] The pattern produced interference of X-rays passing
gives information of the separations of crystallographic planes d, through a crystal. The data can be
allowing one to deduce the crystal structure. Diffraction contrast, in used to determine the crystal's
electron microscopes and x-topography devices in particular, is also atomic structure.
a powerful tool for examining individual defects and local strain
fields in crystals.

Coherence
The description of diffraction relies on the interference of waves emanating from the same source taking
different paths to the same point on a screen. In this description, the difference in phase between waves that
took different paths is only dependent on the effective path length. This does not take into account the fact
that waves that arrive at the screen at the same time were emitted by the source at different times. The initial
phase with which the source emits waves can change over time in an unpredictable way. This means that
waves emitted by the source at times that are too far apart can no longer form a constant interference pattern
since the relation between their phases is no longer time independent.[20]: 9 19 

The length over which the phase in a beam of light is correlated, is called the coherence length. In order for
interference to occur, the path length difference must be smaller than the coherence length. This is
sometimes referred to as spectral coherence, as it is related to the presence of different frequency
components in the wave. In the case of light emitted by an atomic transition, the coherence length is related
to the lifetime of the excited state from which the atom made its transition.[21]: 7 1–74 [22]: 3 14–316 

If waves are emitted from an extended source, this can lead to incoherence in the transversal direction.
When looking at a cross section of a beam of light, the length over which the phase is correlated is called
the transverse coherence length. In the case of Young's double slit experiment, this would mean that if the
transverse coherence length is smaller than the spacing between the two slits, the resulting pattern on a
screen would look like two single slit diffraction patterns.[21]: 7 4–79 

In the case of particles like electrons, neutrons, and atoms, the coherence length is related to the spatial
extent of the wave function that describes the particle.[23]: 1 07 

Applications

Diffraction before destruction

A new way to image single biological particles has emerged over the last few years, utilising the bright X-
rays generated by X-ray free electron lasers. These femtosecond-duration pulses will allow for the
(potential) imaging of single biological macromolecules. Due to these short pulses, radiation damage can be
outrun, and diffraction patterns of single biological macromolecules will be able to be obtained.[24][25]

See also
Angle-sensitive pixel
Atmospheric diffraction
Bragg diffraction
Brocken spectre
Cloud iridescence
Coherent diffraction imaging
Diffraction formalism
Diffraction limit
Diffraction spike
Diffraction vs. interference
Diffractometer
Dynamical theory of diffraction
Electron diffraction
Fraunhofer diffraction
Fresnel diffraction
Fresnel imager
Fresnel number
Fresnel zone
Neutron diffraction
Prism
Powder diffraction
Quasioptics
Refraction
Reflection
Schaefer–Bergmann diffraction
Thinned array curse
X-ray scattering techniques

References
1. Francesco Maria Grimaldi, Physico mathesis de lumine, coloribus, et iride, aliisque annexis
libri duo (Bologna ("Bonomia"), Italy: Vittorio Bonati, 1665), page 2 (https://books.google.co
m/books?id=FzYVAAAAQAAJ&pg=PA2#v=onepage&q&f=false) Archived (https://web.archi
ve.org/web/20161201153749/https://books.google.com/books?id=FzYVAAAAQAAJ&pg=PA
2) 2016-12-01 at the Wayback Machine:

Original  : Nobis alius quartus modus illuxit, quem nunc proponimus,


vocamusque; diffractionem, quia advertimus lumen aliquando diffringi, hoc est
partes eius multiplici dissectione separatas per idem tamen medium in diversa
ulterius procedere, eo modo, quem mox declarabimus.

Translation : It has illuminated for us another, fourth way, which we now make
known and call "diffraction" [i.e., shattering], because we sometimes observe
light break up; that is, that parts of the compound [i.e., the beam of light],
separated by division, advance farther through the medium but in different
[directions], as we will soon show.

2. Cajori, Florian "A History of Physics in its Elementary Branches, including the evolution of
physical laboratories." (https://archive.org/details/ahistoryphysics00cajogoog/page/n102)
Archived (https://web.archive.org/web/20161201075614/https://books.google.com/books?id
=KZ4C-1CRtYQC&ots=c_YpkkbTpT&dq=Florian%20Cajori%20history%20of%20physics&
pg=PA88) 2016-12-01 at the Wayback Machine MacMillan Company, New York 1899
3. Wireless Communications: Principles and Practice, Prentice Hall communications
engineering and emerging technologies series, T. S. Rappaport, Prentice Hall, 2002 pg 126
4. Juffmann, Thomas; Milic, Adriana; Müllneritsch, Michael; Asenbaum, Peter; Tsukernik,
Alexander; Tüxen, Jens; Mayor, Marcel; Cheshnovsky, Ori; Arndt, Markus (2012-03-25).
"Real-time single-molecule imaging of quantum interference". Nature Nanotechnology. 7 (5):
297–300. arXiv:1402.1867 (https://arxiv.org/abs/1402.1867). Bibcode:2012NatNa...7..297J
(https://ui.adsabs.harvard.edu/abs/2012NatNa...7..297J). doi:10.1038/nnano.2012.34 (http
s://doi.org/10.1038%2Fnnano.2012.34). ISSN 1748-3395 (https://www.worldcat.org/issn/174
8-3395). PMID 22447163 (https://pubmed.ncbi.nlm.nih.gov/22447163). S2CID 5918772 (http
s://api.semanticscholar.org/CorpusID:5918772).
5. Francesco Maria Grimaldi, Physico-mathesis de lumine, coloribus, et iride, aliisque adnexis
… [The physical mathematics of light, color, and the rainbow, and other things appended …]
(Bologna ("Bonomia"), (Italy): Vittorio Bonati, 1665), pp. 1–11 (https://books.google.com/boo
ks?id=FzYVAAAAQAAJ&pg=PA1#v=onepage&q&f=false) Archived (https://web.archive.org/
web/20161201074612/https://books.google.com/books?id=FzYVAAAAQAAJ&pg=PA1)
2016-12-01 at the Wayback Machine: "Propositio I. Lumen propagatur seu diffunditur non
solum directe, refracte, ac reflexe, sed etiam alio quodam quarto modo, diffracte."
(Proposition 1. Light propagates or spreads not only in a straight line, by refraction, and by
reflection, but also by a somewhat different fourth way: by diffraction.) On p. 187, Grimaldi
also discusses the interference of light from two sources: "Propositio XXII. Lumen aliquando
per sui communicationem reddit obscuriorem superficiem corporis aliunde, ac prius
illustratam." (Proposition 22. Sometimes light, as a result of its transmission, renders dark a
body's surface, [which had been] previously illuminated by another [source].)
6. Jean Louis Aubert (1760). Memoires pour l'histoire des sciences et des beaux arts (https://ar
chive.org/details/memoirespourlhi146aubegoog). Paris: Impr. de S. A. S.; Chez E. Ganeau.
pp. 149 (https://archive.org/details/memoirespourlhi146aubegoog/page/n151). "grimaldi
diffraction 0-1800."
7. Sir David Brewster (1831). A Treatise on Optics (https://archive.org/details/atreatiseonopti00
brewgoog). London: Longman, Rees, Orme, Brown & Green and John Taylor. pp. 95 (https://
archive.org/details/atreatiseonopti00brewgoog/page/n113).
8. Letter from James Gregory to John Collins, dated 13 May 1673. Reprinted in:
Correspondence of Scientific Men of the Seventeenth Century …, ed. Stephen Jordan
Rigaud (Oxford, England: Oxford University Press, 1841), vol. 2, pp. 251–255, especially p.
254 (https://books.google.com/books?id=0h45L_66bcYC&pg=PA254) Archived (https://web.
archive.org/web/20161201061930/https://books.google.com/books?id=0h45L_66bcYC&pg=
PA254) 2016-12-01 at the Wayback Machine.
9. Thomas Young (1804-01-01). "The Bakerian Lecture: Experiments and calculations relative
to physical optics" (https://books.google.com/books?id=7AZGAAAAMAAJ&pg=PA1).
Philosophical Transactions of the Royal Society of London. 94: 1–16.
Bibcode:1804RSPT...94....1Y (https://ui.adsabs.harvard.edu/abs/1804RSPT...94....1Y).
doi:10.1098/rstl.1804.0001 (https://doi.org/10.1098%2Frstl.1804.0001). S2CID 110408369
(https://api.semanticscholar.org/CorpusID:110408369).. (Note: This lecture was presented
before the Royal Society on 24 November 1803.)
10. Fresnel, Augustin-Jean (1816), "Mémoire sur la diffraction de la lumière" ("Memoir on the
diffraction of light"), Annales de Chimie et de Physique, vol. 1, pp. 239–81 (March 1816);
reprinted as "Deuxième Mémoire…" ("Second Memoir…") in Oeuvres complètes d'Augustin
Fresnel, vol. 1 (Paris: Imprimerie Impériale, 1866), pp. 89–122 (https://books.google.com/bo
oks?id=1l0_AAAAcAAJ&pg=PA89). (Revision of the "First Memoir" (https://books.google.co
m/books?id=1l0_AAAAcAAJ&pg=PA9) submitted on 15 October 1815.)
11. Fresnel, Augustin-Jean (1818), "Mémoire sur la diffraction de la lumière" ("Memoir on the
diffraction of light"), deposited 29 July 1818, "crowned" 15 March 1819, published in
Mémoires de l'Académie Royale des Sciences de l'Institut de France, vol. V (for 1821 &
1822, printed 1826), pp. 339–475 (https://books.google.com/books?id=zNo-AQAAMAAJ&pg
=PA339); reprinted in Oeuvres complètes d'Augustin Fresnel, vol. 1 (Paris: Imprimerie
Impériale, 1866), pp. 247–364 (https://books.google.com/books?id=1l0_AAAAcAAJ&pg=PA
247); partly translated as "Fresnel's prize memoir on the diffraction of light" (https://archive.or
g/details/wavetheoryofligh00crewrich/page/80), in H. Crew (ed.), The Wave Theory of Light:
Memoirs by Huygens, Young and Fresnel, American Book Company, 1900, pp. 81–144.
(First published, as extracts only, in Annales de Chimie et de Physique, vol. 11 (1819),
pp. 246–96 (https://books.google.com/books?id=SSRQAAAAcAAJ&pg=PA246), 337–78 (htt
ps://books.google.com/books?id=SSRQAAAAcAAJ&pg=PA337).)
12. Christiaan Huygens, Traité de la lumiere … (https://archive.org/details/bub_gb_X9PKaZlCh
ggC) Archived (https://web.archive.org/web/20160616191659/https://books.google.com/boo
ks?id=X9PKaZlChggC&pg=PP5) 2016-06-16 at the Wayback Machine (Leiden,
Netherlands: Pieter van der Aa, 1690), Chapter 1. From p. 15 (https://archive.org/details/bub
_gb_X9PKaZlChggC/page/n94) Archived (https://web.archive.org/web/20161201064555/htt
ps://books.google.com/books?id=X9PKaZlChggC&pg=PA15) 2016-12-01 at the Wayback
Machine: "J'ay donc monstré de quelle façon l'on peut concevoir que la lumiere s'etend
successivement par des ondes spheriques, … " (I have thus shown in what manner one can
imagine that light propagates successively by spherical waves, … ) (Note: Huygens
published his Traité in 1690; however, in the preface to his book, Huygens states that in
1678 he first communicated his book to the French Royal Academy of Sciences.)
13. Baker, B.B. & Copson, E.T. (1939), The Mathematical Theory of Huygens' Principle, Oxford,
pp. 36–40.
14. Dietrich Zawischa. "Optical effects on spider webs" (http://www.itp.uni-hannover.de/%7Ezaw
ischa/ITP/spiderweb.html). Retrieved 2007-09-21.
15. Arumugam, Nadia. "Food Explainer: Why Is Some Deli Meat Iridescent?" (http://www.slate.c
om/blogs/browbeat/2013/09/09/iridescent_deli_meat_why_some_sliced_ham_and_beef_sh
ine_with_rainbow_colors.html). Slate. The Slate Group. Archived (https://web.archive.org/we
b/20130910021203/http://www.slate.com/blogs/browbeat/2013/09/09/iridescent_deli_meat_
why_some_sliced_ham_and_beef_shine_with_rainbow_colors.html) from the original on 10
September 2013. Retrieved 9 September 2013.
16. Andrew Norton (2000). Dynamic fields and waves of physics (https://books.google.com/book
s?id=XRRMxjr24pwC&pg=PA102). CRC Press. p. 102. ISBN 978-0-7503-0719-2.
17. Chiao, R. Y.; Garmire, E.; Townes, C. H. (1964). "Self-Trapping of Optical Beams" (http://jour
nals.aps.org/prl/pdf/10.1103/PhysRevLett.13.479). Physical Review Letters. 13 (15): 479–
482. Bibcode:1964PhRvL..13..479C (https://ui.adsabs.harvard.edu/abs/1964PhRvL..13..479
C). doi:10.1103/PhysRevLett.13.479 (https://doi.org/10.1103%2FPhysRevLett.13.479).
18. Brezger, B.; Hackermüller, L.; Uttenthaler, S.; Petschinka, J.; Arndt, M.; Zeilinger, A.
(February 2002). "Matter–Wave Interferometer for Large Molecules" (http://homepage.univie.
ac.at/Lucia.Hackermueller/unsereArtikel/Brezger2002a.pdf) (reprint). Physical Review
Letters. 88 (10): 100404. arXiv:quant-ph/0202158 (https://arxiv.org/abs/quant-ph/0202158).
Bibcode:2002PhRvL..88j0404B (https://ui.adsabs.harvard.edu/abs/2002PhRvL..88j0404B).
doi:10.1103/PhysRevLett.88.100404 (https://doi.org/10.1103%2FPhysRevLett.88.100404).
PMID 11909334 (https://pubmed.ncbi.nlm.nih.gov/11909334). S2CID 19793304 (https://api.s
emanticscholar.org/CorpusID:19793304). Archived (https://web.archive.org/web/200708130
92642/http://homepage.univie.ac.at/Lucia.Hackermueller/unsereArtikel/Brezger2002a.pdf)
(PDF) from the original on 2007-08-13. Retrieved 2007-04-30.
19. John M. Cowley (1975) Diffraction physics (North-Holland, Amsterdam) ISBN 0-444-10791-
6
20. Halliday, David; Resnick, Robert; Walker, Jerl (2005), Fundamental of Physics (https://archiv
e.org/details/isbn_0471216437) (7th ed.), USA: John Wiley and Sons, Inc., ISBN 978-0-471-
23231-5
21. Grant R. Fowles (1975). Introduction to Modern Optics. Courier Corporation. ISBN 978-0-
486-65957-2.
22. Hecht, Eugene (2002), Optics (4th ed.), United States of America: Addison Wesley,
ISBN 978-0-8053-8566-3
23. Ayahiko Ichimiya; Philip I. Cohen (13 December 2004). Reflection High-Energy Electron
Diffraction (https://books.google.com/books?id=AUVbPerNxTcC). Cambridge University
Press. ISBN 978-0-521-45373-8. Archived (https://web.archive.org/web/20170716041343/ht
tps://books.google.com/books?id=AUVbPerNxTcC) from the original on 16 July 2017.
24. Neutze, Richard; Wouts, Remco; van der Spoel, David; Weckert, Edgar; Hajdu, Janos
(August 2000). "Potential for biomolecular imaging with femtosecond X-ray pulses" (https://w
ww.nature.com/articles/35021099). Nature. 406 (6797): 752–757. doi:10.1038/35021099 (htt
ps://doi.org/10.1038%2F35021099). ISSN 1476-4687 (https://www.worldcat.org/issn/1476-4
687). PMID 10963603 (https://pubmed.ncbi.nlm.nih.gov/10963603). S2CID 4300920 (https://
api.semanticscholar.org/CorpusID:4300920).
25. Chapman, Henry N.; Caleman, Carl; Timneanu, Nicusor (2014-07-17). "Diffraction before
destruction" (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC4052855). Philosophical
Transactions of the Royal Society B: Biological Sciences. 369 (1647): 20130313.
doi:10.1098/rstb.2013.0313 (https://doi.org/10.1098%2Frstb.2013.0313). PMC 4052855 (http
s://www.ncbi.nlm.nih.gov/pmc/articles/PMC4052855). PMID 24914146 (https://pubmed.ncbi.
nlm.nih.gov/24914146).

External links
"Scattering and diffraction" (https://www.xtal.iqfr.csic.es/Cristalografia/parte_05-en.html).
Crystallography. International Union of Crystallography.

Retrieved from "https://en.wikipedia.org/w/index.php?title=Diffraction&oldid=1067715213"

This page was last edited on 24 January 2022, at 20:22 (UTC).

Text is available under the Creative Commons Attribution-ShareAlike License 3.0;


additional terms may apply. By
using this site, you agree to the Terms of Use and Privacy Policy. Wikipedia® is a registered trademark of the
Wikimedia Foundation, Inc., a non-profit organization.

You might also like