Presentation is loading. Please wait.

Presentation is loading. Please wait.

FURKAN KAMIL MOHAMMAD.

Similar presentations


Presentation on theme: "FURKAN KAMIL MOHAMMAD."— Presentation transcript:

1 FURKAN KAMIL MOHAMMAD

2 From Wikipedia, the free encyclopedia
 Diffraction From Wikipedia, the free encyclopedia Computer generated intensity pattern formed on a screen by diffraction from a square aperture

3 Generation of an interference pattern from two-slit diffraction

4 Computational Model of an interference patern from two-slit diffraction

5

6 Colors seen in a spider web are partially due to diffraction, according to some analyses.[1]
Diffraction refers to various phenomena which occur when a wave encounters an obstacle. Italian scientist Francesco Maria Grimaldi coined the word "diffraction" and was the first to record accurate observations of the phenomenon in 1665.[2][3] In classical physics, the diffraction phenomenon is described as the apparent bending of waves around small obstacles and the spreading out of waves past small openings. Similar effects occur when light waves travel through a medium with a varying refractive index or a sound wave through one with varying acoustic impedance. Diffraction occurs with all waves, including sound waves, water waves, and electromagnetic waves such as visible light, x-rays and radio waves. As physical objects have wave-like properties (at the atomic level), diffraction also occurs with matter and can be studied according to the principles of quantum mechanics. Richard Feynman[4] said that "no-one has ever been able to define the difference between interference and diffraction satisfactorily. It is just a question of usage, and there is no specific, important physical difference between them." He suggested that when there are only a few sources, say two, we call it interference, as in Young's slits, but with a large number of sources, the process is labelled diffraction. While diffraction occurs whenever propagating waves encounter such changes, its effects are generally most pronounced for waves where the wavelength is roughly similar to the dimensions of the diffracting objects. If the obstructing object provides multiple, closely spaced openings, a complex pattern of varying intensity can result. This is due to the superposition, or interference, of different parts of a wave that traveled to the observer by different paths (see diffraction grating). The formalism of diffraction can also describe the way in which waves of finite extent propagate in free space. For example, the expanding profile of a laser beam, the beam shape of a radar antenna and the field of view of an ultrasonic transducer can all be analysed using diffraction equations.

7 Examples

8 Solar glory at the steam from hot springs
Solar glory at the steam from hot springs. A glory is an optical phenomenon produced by light backscattered (a combination of diffraction, reflection and refraction) towards its source by a cloud of uniformly sized water droplets. The effects of diffraction are often seen in everyday life. The most striking examples of diffraction are those involving 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 disk. 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 source, shows small fringes near its edges. The speckle pattern which is observed when laser light falls on an optically rough surface is also a diffraction phenomenon. All these effects are a consequence of the fact that light propagates as a wave.

9 Diffraction can occur with any kind of 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.[5] Diffraction can also be a concern in some technical applications; it sets a fundamental limit to the resolution of a camera, telescope, or microscope. History Thomas Young's sketch of two-slit diffraction, which he presented to the Royal Society in 1803 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.[6][7][8] Isaac Newton studied these effects and attributed them to inflexion of light rays. James Gregory (1638–1675) observed the diffraction patterns caused by a bird feather, which was effectively the first diffraction grating to be discovered.[9] Thomas Young performed a celebrated experiment in 1803 demonstrating interference from two closely spaced slits.[10] 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 1815[11] and 1818,[12] and thereby gave great support to the wave theory of light that had been advanced by Christiaan Huygens[13] and reinvigorated by Young, against Newton's particle theory.

10 Mechanism Photograph of single-slit diffraction in a circular ripple tank 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 point on a wavefront as

11 [edit] Diffraction of light
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 phases as well as the amplitudes of the individual waves so that the 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. There are various analytical models which allow the diffracted field to be calculated, including the Kirchhoff-Fresnel diffraction equation which is derived from wave equation, 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. [edit] Diffraction of light Some examples of diffraction of light are considered below. [edit] Single-slit diffraction Main article: Diffraction formalism

12 Numerical approximation of diffraction pattern from a slit of width equal to wavelength of an incident plane wave in 3D spectrum visualization

13 Numerical approximation of diffraction pattern from a slit of width equal to five times the wavelength of an incident plane wave in 3D spectrum visualization

14 Diffraction of red laser beam on the hole
Numerical approximation of diffraction pattern from a slit of width four wavelengths with an incident plane wave. The main central beam, nulls, and phase reversals are apparent.

15 so that the minimum intensity occurs at an angle θmin given by
where d is the width of the slit, θmin is the angle of incidence at which the minimum intensity occurs, and λ is the wavelength of the light

16 Graph and image of 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. A slit which is wider than a wavelength produces interference effects in the space downstream of the slit. These can be explained by assuming that the slit behaves as though it has a large number of point sources spaced evenly across the width of the slit. The analysis of this system is simplified if we consider light of a single wavelength. If the incident light is monochromatic, 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 top of the slit will interfere destructively with the source located just 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 given by

17 [edit] Diffraction grating
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 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 I(θ) is the intensity at a given angle, I0 is the original intensity, and the sinc function is given by sinc(x) = sin(πx)/(πx) if x ≠ 0, and sinc(0) = 1 This analysis applies only to the far field, that is, at a distance much larger than the width of the slit. 2-slit (top) and 5-slit diffraction of red laser light Diffraction of a red laser using a diffraction grating A diffraction pattern of a 633 nm laser through a grid of 150 slits [edit] Diffraction grating

18 [edit] Circular aperture
Main article: 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 the grating equation 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. A computer-generated image of an Airy disk Computer generated light diffraction pattern from a circular aperture of diameter 0.5 micrometre at a wavelength of 0.6 micrometre (red-light) at distances of 0.1 cm – 1 cm 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. [edit] Circular aperture

19 [edit] Circular aperture
Main article: 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 the grating equation 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. A computer-generated image of an Airy disk Computer generated light diffraction pattern from a circular aperture of diameter 0.5 micrometre at a wavelength of 0.6 micrometre (red-light) at distances of 0.1 cm – 1 cm 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. [edit] Circular aperture

20 [edit] General aperture
Main article: Airy disk 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 where a is the radius of the circular aperture, k is equal to 2π/λ and 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. [edit] General aperture The wave that emerges from a point source has amplitude ψ 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 (aka 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 e − iωt) is: This solution assumes that the delta function source is located at the origin. If the source is located at an arbitrary source point, denoted by the vector and the field point is located at the point , then we may represent the scalar Green's function (for arbitrary source location) as: 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:

21 On the calculation of Fraunhofer region fields
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, simplifies to Now, since as can be seen in the figure to the right (click to enlarge). The expression for the far-zone (Fraunhofer region) field becomes and the expression for the Fraunhofer region field from a planar aperture now becomes,

22 [edit] Propagation of a laser beam
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). [edit] Propagation of a laser beam The way in which the profile of a laser beam changes as it propagates is determined by diffraction. The output mirror of the laser is an aperture, and the subsequent beam shape is determined by that aperture. Hence, the smaller the output beam, the quicker it diverges. Diode lasers have much greater divergence than He–Ne lasers for this reason. Paradoxically, 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 aperture, and hence a lower divergence. [edit] Diffraction-limited imaging Main article: Diffraction-limited system

23 [edit] Speckle patterns
The Airy disk around each of the stars from the 2.56 m telescope aperture can be seen in this lucky image of the binary star zeta Boötis. 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 with radius to first null of where λ is the wavelength of the light and N is the f-number (focal length divided by diameter) of the imaging optics. In object space, the corresponding angular resolution is 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 can be considered to be resolvable 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, and the smaller the wavelength, the finer the resolution of an imaging system. This is why telescopes have very large lenses or mirrors, and why optical microscopes are limited in the detail which they can see. [edit] Speckle patterns Main article: speckle pattern The speckle pattern which is seen when using a laser pointer is another diffraction phenomenon. It is a result of the superpostion 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.

24 [edit] Particle diffraction
Patterns The upper half of this image shows a diffraction pattern of He-Ne laser beam on an elliptic aperture. The lower half is its 2D Fourier transform approximately reconstructing the shape of the aperture. 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. [edit] Particle diffraction See also: neutron diffraction and electron diffraction Quantum theory tells us that every particle exhibits wave properties. In particular, massive particles can interfere 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

25 [edit] Bragg diffraction
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.[14] [edit] Bragg diffraction where λ is the wavelength, d is the distance between crystal planes, θ is the angle of the diffracted wave. Following Bragg's law, each dot (or reflection), in this diffraction pattern forms from the constructive interference of X-rays passing through a crystal. The data can be used to determine the crystal's atomic structure. For more details on this topic, see Bragg diffraction. Diffraction from a three dimensional periodic structure such as atoms in a crystal is called Bragg diffraction. It is similar to what occurs when waves are scattered from a diffraction grating. Bragg diffraction is a consequence of interference between waves reflecting from different crystal planes. The condition of constructive interference is given by Bragg's law:

26 Fresnel diffraction [edit] Coherence
From Wikipedia, the free encyclopedia and m is an integer known as the order of the diffracted beam. Bragg diffraction may be carried out using either light of very short wavelength like x-rays or matter waves like neutrons (and electrons) whose wavelength is on the order of (or much smaller than) the atomic spacing.[15] The pattern produced gives information of the separations of crystallographic planes d, allowing one to deduce the crystal structure. Diffraction contrast, in electron microscopes and x-topography devices in particular, is also a powerful tool for examining individual defects and local strain fields in crystals. [edit] Coherence Main article: Coherence (physics) 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. 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. 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. 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.

27 [edit] Early treatments of this phenomenon
In optics, the Fresnel diffraction equation for near-field diffraction, is an approximation of Kirchhoff-Fresnel diffraction that can be applied to the propagation of waves in the near field.[1] The near field can be specified by the Fresnel number, F of the optical arrangement, which is defined, for a wave incident on an aperture, as: is the characteristic size of the aperture where is the distance of the observation point from the aperture is the wavelength of the wave. When the diffracted wave is considered to be in the near field, and the Fresnel diffraction equation can be used to calculate its form. Fresnel diffraction showing central Arago spot The multiple Fresnel diffraction at nearly placed periodical ridges (ridged mirror) causes the specular reflection; this effect can be used for atomic mirrors.[2] [edit] Early treatments of this phenomenon Some of the earliest work on what would become known as Fresnel diffraction was carried out by Francesco Maria Grimaldi in 17th century Italy. In his monograph entitled "Light,"[3] Richard C. MacLaurin explains Fresnel diffraction by asking what happens when light propagates, and how that process is affected when a barrier with a slit or hole in it is interposed in the beam produced by a distant source of light. He uses the Principle of Huygens to investigate, in classical terms, what transpires. The wave front that proceeds from the slit and on to a detection screen some distance away very closely approximates a wave front originating across the area of the gap without regard to any minute interactions with the actual physical edge.

28 The result is that if the gap is very narrow only diffraction patterns with bright centers can occur. If the gap is made progressively wider, then diffraction patterns with dark centers will alternate with diffraction patterns with bright centers. As the gap becomes larger, the differentials between dark and light bands decrease until a diffraction effect can no longer be detected. MacLaurin does not mention the possibility that the center of the series of diffraction rings produced when light is shone through a small hole may be black, but he does point to the inverse situation wherein the shadow produced by a small circular object can paradoxically have a bright center. (p. 219) In his Optics,,[4] Francis Weston Sears offers a mathematical approximation suggested by Fresnel that predicts the main features of diffraction patterns and uses only simple mathematics. By considering the perpendicular distance from the hole in a barrier screen to a nearby detection screen along with the wavelength of the incident light, it is possible to compute a number of regions called half-period elements or Fresnel zones. The inner zone will be a circle and each succeeding zone will be a concentric annular ring. If the diameter of the circular hole in the screen is sufficient to expose the first or central Fresnel zone, the amplitude of light at the center of the detection screen will be double what it would be if the detection screen were not obstructed. If the diameter of the circular hole in the screen is sufficient to expose two Fresnel zones, then the amplitude at the center is almost zero. That means that a Fresnel diffraction pattern can have a dark center. These patterns can be seen and measured, and correspond well to the values calculated for them. Figure 9-5, following p. 222, in Sears shows four Fraunhofer patterns in the top row followed by sixteen Fresnel diffraction patterns. Three of them have dark centers. (See the photograph above, and check the book by Sears for very much nicer photographs. A simulation can be operated by visiting

29 The Fresnel diffraction integral
Diffraction geometry, showing aperture (or diffracting object) plane and image plane, with coordinate system. The electric field diffraction pattern at a point (x,y,z) is given by:

30 Substituting into the expression for r, we find:
where , and Substituting into the expression for r, we find: is the imaginary unit. Analytical solution of this integral is impossible for all but the simplest diffraction geometries. Therefore, it is usually calculated numerically. [edit] The Fresnel approximation The main problem for solving the integral is the expression of r. First, we can simplify the algebra by introducing the substitution: Next, using the Taylor series expansion we can express r as If we consider all the terms of Taylor series, then there is no approximation.[5] Let us substitute this expression in the argument of the exponential within the integral; the key to the Fresnel approximation is to assume that the third element is very small and can be ignored. In order to make this possible, it has to contribute to the variation of the exponential for an almost null term. In other words, it has to be much smaller than the period of the complex exponential, i.e. 2π:

31 expressing k in terms of the wavelength,
we get the following relationship: Multiplying both sides by z3 / λ3, we have or, substituting the earlier expression for ρ2 , If this condition holds true for all values of x, x' , y and y' , then we can ignore the third term in the Taylor expression. Furthermore, if the third term is negligible, then all terms of higher order will be even smaller, so we can ignore them as well. For applications involving optical wavelengths, the wavelength λ is typically many orders of magnitude smaller than the relevant physical dimensions. In particular: and Thus, as a practical matter, the required inequality will always hold true as long as We can then approximate the expression with only the first two terms: This equation, then, is the Fresnel approximation, and the inequality stated above is a condition for the approximation's validity.

32 [edit] Alternative forms
[edit] Fresnel diffraction The condition for validity is fairly weak, and it allows all length parameters to take comparable values, provided the aperture is small compared to the path length. For the r in the denominator we go one step further, and approximate it with only the first term, . This is valid in particular if we are interested in the behaviour of the field only in a small area close to the origin, where the values of x and y are much smaller than z. In addition, it is always valid if as well as the Fresnel condition, we have , where L is the distance between the aperture and the field point. For Fresnel diffraction the electric field at point (x,y,z) is then given by: This is the Fresnel diffraction integral; it means that, if the Fresnel approximation is valid, the propagating field is a spherical wave, originating at the aperture and moving along z. The integral modulates the amplitude and phase of the spherical wave. Analytical solution of this expression is still only possible in rare cases. For a further simplified case, valid only for much larger distances from the diffraction source see Fraunhofer diffraction. Unlike Fraunhofer diffraction, Fresnel diffraction accounts for the curvature of the wavefront, in order to correctly calculate the relative phase of interfering waves. [edit] Alternative forms [edit] Convolution The integral can be expressed in other ways in order to calculate it using some mathematical properties. If we define the following function: then the integral can be expressed in terms of a convolution: in other words we are representing the propagation using a linear-filter modeling. That is why we might call the function h(x,y,z) the impulse response of free space propagation. [edit] Fourier transform Fresnel diffraction occurs when: Fraunhofer diffraction occurs when: a - aperture or slit size, λ - wavelength, L - distance from the aperture

33 Fresnel integral [edit] Linear canonical transformation
Another possible way is through the Fourier transform. If in the integral we express k in terms of the wavelength, and we expand each component of the transverse displacement, then we can express the integral in terms of the two dimensional Fourier transform. Let us use the following definition: where p and q are spatial frequencies (in units of lines/meter). The Fresnel integral can be expressed as: i.e. first multiply the field to be propagated for a complex exponential, calculate its two dimensional Fourier transform, replace (p,q) with Fresnel integral and multiply it by another factor. This expression is better than the others when the process leads to a known Fourier transform, and the connection with the Fourier transform is tightened in the linear canonical transformation, discussed below. [edit] Linear canonical transformation Main article: Linear canonical transformation From the point of view of the linear canonical transformation, Fresnel diffraction can be seen as a shear in the time-frequency domain, corresponding to how the Fourier transform is a rotation in the time-frequency domain.

34 [edit] Definition From Wikipedia, the free encyclopedia
S(x) and C(x) The maximum of C(x) is about If πt²/2 were used instead of t², then the image would be scaled vertically and horizontally (see below). Fresnel integrals, S(x) and C(x), are two transcendental functions named after Augustin-Jean Fresnel that are used in optics. They arise in the description of near field Fresnel diffraction phenomena, and are defined through the following integral representations: The simultaneous parametric plot of S(x) and C(x) is the Euler spiral, also known as the Cornu spiral or clothoid. [edit] Definition The Fresnel integrals admit the following power series expansions that converge for all x: Normalised Fresnel integrals, S(x) and C(x). In these curves, the argument of the trigonometric function is πt2/2, as opposed to just t2 as above.

35 Some authors, including Abramowitz and Stegun, (eqs 7.3.1 – 7.3.2) use
for the argument of the integrals defining S(x) and C(x). To get these functions, multiply the above integrals by and divide the argument x by the same factor. [edit] Euler spiral Main article: Euler spiral Euler spiral (x, y) = (C(t), S(t)). The spiral converges to the centre of the holes in the image as t tends to positive or negative infinity. The Euler spiral, also known as Cornu spiral or clothoid, is the curve generated by a parametric plot of S(t) against C(t). The Cornu spiral was created by Marie Alfred Cornu as a nomogram for diffraction computations in science and engineering. From the definitions of Fresnel integrals, the infinitesimals dx and dy are thus: Thus the length of the spiral measured from the origin can be expressed as: That is, the parameter t is the curve length measured from the origin (0,0) and the Euler spiral has infinite length. The vector [cos(t²), sin(t²)] also expresses the unit tangent vector along the spiral, giving θ = t². Since t is the curve length, the curvature, κ can be expressed as:

36 Some authors, including Abramowitz and Stegun, (eqs 7.3.1 – 7.3.2) use
for the argument of the integrals defining S(x) and C(x). To get these functions, multiply the above integrals by and divide the argument x by the same factor. [edit] Euler spiral Main article: Euler spiral Euler spiral (x, y) = (C(t), S(t)). The spiral converges to the centre of the holes in the image as t tends to positive or negative infinity. The Euler spiral, also known as Cornu spiral or clothoid, is the curve generated by a parametric plot of S(t) against C(t). The Cornu spiral was created by Marie Alfred Cornu as a nomogram for diffraction computations in science and engineering. From the definitions of Fresnel integrals, the infinitesimals dx and dy are thus: Thus the length of the spiral measured from the origin can be expressed as: That is, the parameter t is the curve length measured from the origin (0,0) and the Euler spiral has infinite length. The vector [cos(t²), sin(t²)] also expresses the unit tangent vector along the spiral, giving θ = t². Since t is the curve length, the curvature, κ can be expressed as:

37 [edit] Euler spiral Main article: Euler spiral
Euler spiral (x, y) = (C(t), S(t)). The spiral converges to the centre of the holes in the image as t tends to positive or negative infinity. The Euler spiral, also known as Cornu spiral or clothoid, is the curve generated by a parametric plot of S(t) against C(t). The Cornu spiral was created by Marie Alfred Cornu as a nomogram for diffraction computations in science and engineering. From the definitions of Fresnel integrals, the infinitesimals dx and dy are thus: Thus the length of the spiral measured from the origin can be expressed as: That is, the parameter t is the curve length measured from the origin (0,0) and the Euler spiral has infinite length. The vector [cos(t²), sin(t²)] also expresses the unit tangent vector along the spiral, giving θ = t². Since t is the curve length, the curvature, κ can be expressed as:

38 Fresnel integral [edit] Generalization
The sector contour used to calculate the limits of the Fresnel integrals The limits of C and S as the argument tends to infinity can be found by the methods of complex analysis. This uses the contour integral of the function around the boundary of the sector-shaped region in the complex plane formed by the positive x-axis, the half-line y = x, x ≥ 0, and the circle of radius R centered at the origin. As R goes to infinity, the integral along the circular arc tends to 0, the integral along the real axis tends to the Gaussian integral and after routine transformations, the integral along the bisector of the first quadrant can be related to the limit of the Fresnel integrals. [edit] Generalization The Fresnel integral can be generalized by the function with the left-hand side converging for a>1 and the right-hand side being its analytical extension to the whole plane less where lie the poles of Γ(a − 1). Fresnel integral From Wikipedia, the free encyclopedia

39 S(x) and C(x) The maximum of C(x) is about 0. 977451424
S(x) and C(x) The maximum of C(x) is about If πt²/2 were used instead of t², then the image would be scaled vertically and horizontally (see below). Fresnel integrals, S(x) and C(x), are two transcendental functions named after Augustin-Jean Fresnel that are used in optics. They arise in the description of near field Fresnel diffraction phenomena, and are defined through the following integral representations: The simultaneous parametric plot of S(x) and C(x) is the Euler spiral, also known as the Cornu spiral or clothoid. [edit] Definition The Fresnel integrals admit the following power series expansions that converge for all x: Normalised Fresnel integrals, S(x) and C(x). In these curves, the argument of the trigonometric function is πt2/2, as opposed to just t2 as above.

40 Some authors, including Abramowitz and Stegun, (eqs 7.3.1 – 7.3.2) use
for the argument of the integrals defining S(x) and C(x). To get these functions, multiply the above integrals by and divide the argument x by the same factor. [edit] Euler spiral Main article: Euler spiral Euler spiral (x, y) = (C(t), S(t)). The spiral converges to the centre of the holes in the image as t tends to positive or negative infinity. The Euler spiral, also known as Cornu spiral or clothoid, is the curve generated by a parametric plot of S(t) against C(t). The Cornu spiral was created by Marie Alfred Cornu as a nomogram for diffraction computations in science and engineering. From the definitions of Fresnel integrals, the infinitesimals dx and dy are thus: Thus the length of the spiral measured from the origin can be expressed as: That is, the parameter t is the curve length measured from the origin (0,0) and the Euler spiral has infinite length. The vector [cos(t²), sin(t²)] also expresses the unit tangent vector along the spiral, giving θ = t². Since t is the curve length, the curvature, κ can be expressed as:

41 [edit] Properties [edit] Evaluation
And the rate of change of curvature with respect to the curve length is: An Euler spiral has the property that its curvature at any point is proportional to the distance along the spiral, measured from the origin. This property makes it useful as a transition curve in highway and railway engineering. If a vehicle follows the spiral at unit speed, the parameter t in the above derivatives also represents the time. That is, a vehicle following the spiral at constant speed will have a constant rate of angular acceleration. Sections from Euler spirals are commonly incorporated into the shape of roller-coaster loops to make what are known as "clothoid loops". [edit] Properties C(x) and S(x) are odd functions of x. C and S are entire functions. Using the power series expansions above, the Fresnel integrals can be extended to the domain of complex numbers, and they become analytic functions of a complex variable. The Fresnel integrals can be expressed using the error function as follows: The integrals defining C(x) and S(x) cannot be evaluated in the closed form in terms of elementary functions, except in special cases. The limits of these functions as x goes to infinity are known: [edit] Evaluation

42 Diffraction grating [edit] Generalization
The sector contour used to calculate the limits of the Fresnel integrals The limits of C and S as the argument tends to infinity can be found by the methods of complex analysis. This uses the contour integral of the function around the boundary of the sector-shaped region in the complex plane formed by the positive x-axis, the half-line y = x, x ≥ 0, and the circle of radius R centered at the origin. As R goes to infinity, the integral along the circular arc tends to 0, the integral along the real axis tends to the Gaussian integral and after routine transformations, the integral along the bisector of the first quadrant can be related to the limit of the Fresnel integrals. [edit] Generalization The Fresnel integral can be generalized by the function with the left-hand side converging for a>1 and the right-hand side being its analytical extension to the whole plane less where lie the poles of Γ(a − 1). Diffraction grating A very large reflecting diffraction grating.

43 In optics, a diffraction grating is an optical component with a periodic structure, which splits and diffracts light into several beams travelling in different directions. The directions of these beams depend on the spacing of the grating and the wavelength of the light so that the grating acts as the dispersive element. Because of this, gratings are commonly used in monochromators and spectrometers. A photographic slide with a fine pattern of purple lines forms a complex grating. For practical applications, gratings generally have ridges or rulings on their surface rather than dark lines. Such gratings can be either transmissive or reflective. Gratings which modulate the phase rather than the amplitude of the incident light are also produced, frequently using holography. The principles of diffraction gratings were discovered by James Gregory, about a year after Newton's prism experiments, initially with artifacts such as bird feathers. The first man-made diffraction grating was made around 1785 by Philadelphia inventor David Rittenhouse, who strung hairs between two finely threaded screws. This was similar to notable German physicist Joseph von Fraunhofer's wire diffraction grating in 1821. of operation Main article: Diffraction A diffraction grating reflecting green light from a dye laser. The relationship between the grating spacing and the angles of the incident and diffracted beams of light is known as the grating equation. According to the Huygens–Fresnel principle, each point on the wavefront of a propagating wave can be considered to act as a point source, and the wavefront at any subsequent point can be found by adding together the contributions from each of these individual point sources. An idealised grating is considered here which is made up of a set of long and infinitely narrow slits of spacing d. When a plane wave of wavelength λ is incident normally on the grating, each slit in the grating acts as a point source propagating in all directions. The light in a particular direction, θ, is made up of the interfering

44 components from each slit
components from each slit. Generally, the phases of the waves from different slits will vary from one another, and will cancel one another out partially or wholly. However, when the path difference between the light from adjacent slits is equal to the wavelength, λ, the waves will all be in phase. This occurs at angles θm which satisfy the relationship dsinθm/λ=|m| where d is the separation of the slits and m is an integer. Thus, the diffracted light will have maxima at angles θm given by It is straightforward to show that if a plane wave is incident at an angle θi, the grating equation becomes The light that corresponds to direct transmission (or specular reflection in the case of a reflection grating) is called the zero order, and is denoted m = 0. The other maxima occur at angles which are represented by non-zero integers m. Note that m can be positive or negative, resulting in diffracted orders on both sides of the zero order beam. This derivation of the grating equation has used an idealised grating. However, the relationship between the angles of the diffracted beams, the grating spacing and the wavelength of the light apply to any regular structure of the same spacing, because the phase relationship between light scattered from adjacent elements of the grating remains the same. The detailed distribution of the diffracted light depends on the detailed structure of the grating elements as well as on the number of elements in the grating, but it will always give maxima in the directions given by the grating equation. Gratings can be made in which various properties of the incident light are modulated in a regular pattern; these include transparency (transmission amplitude gratings) reflectance (reflection amplitude gratings) refractive index (phase gratings) direction of optical axis (optical axis gratings) The grating equation applies in all these cases. [edit] QED QED (quantum electrodynamics) offers another derivation of the properties of a diffraction grating in terms of photons as particles. QED models photons as following all paths from a source to a final point, each of which has a certain probability amplitude, which can be represented as a vector or complex number (equivalently), or as Richard Feynman simply calls them in his book on QED, "arrows". For the probability that a certain event will happen, one sums the probability amplitudes for all of the possible ways in which the event can occur, and then takes the square of the length of the result. The probability amplitude of a photon from a monochromatic source, in this case, is modeled as an arrow that spins rapidly until it

45 [edit] Gratings as dispersive elements
is "evaluated" when the photon reaches its final point. This spinning is actually dependent on the time at which the photon would have left the monochromatic source, as the probability amplitudes of photons do not "spin" while they are in transit. For example, for the probability that light will reflect off of a mirror, one sets the photon's probability amplitude spinning as it leaves the source, follows it to the mirror, and then to its final point,even for paths that do not involve bouncing off of the mirror at equal angles. One then "evaluates" the probability amplitude at the photon's final point; next, one sums these arrows in a standard vector sum, and squares the length of the result for the probability that this photon will reflect off of the mirror. The times these paths take are what determine the angle of the probability amplitude arrow, as they 'spin' at a constant rate (which is related to the frequency of the photon). The times of the paths near the classical reflection site of the mirror will be nearly the same, so as a result the probability amplitudes will point in nearly the same direction—thus, they will have a sizable sum. Examining the paths towards the edges of the mirror reveals that the times of nearby paths are quite different from each other, and thus we wind up summing vectors that cancel out quickly. So, there is a higher probability that light will follow a near-classical reflection path than a path further out. However, a diffraction grating can be made out of this mirror, by scraping away areas near the edge of the mirror that usually cancel nearby amplitudes out—but now, since the photons would not reflect from the scraped-off portions, the probability amplitudes which all point, for instance, to the right can have a sizable sum. Thus, this would let light of the right frequency sum to a larger probability amplitude, and as such possess a larger probability of reflecting. This particular description involves many simplifications: a point source, a "surface" that light can reflect off of (thus neglecting the interactions with electrons) and so forth. However, this approximation is a reasonable one to illustrate a diffraction grating conceptually. Light of a different frequency can also use the same diffraction grating, but with a different final point.[1] [edit] Gratings as dispersive elements The wavelength dependence in the grating equation shows that the grating separates an incident polychromatic beam into its constituent wavelength components, i.e., it is dispersive. Each wavelength of input beam spectrum is sent into a different direction, producing a rainbow of colors under white light illumination. This is visually similar to the operation of a prism, although the mechanism is very different. A light bulb of a flashlight seen through a transmissive grating, showing three diffracted orders. The order m = 0 corresponds to a direct transmission of ligh

46 through the grating. In the first positive order (m = +1), colors with increasing wavelengths (from blue to red) are diffracted at increasing angles. The diffracted beams corresponding to consecutive orders may overlap, depending on the spectral content of the incident beam and the grating density. The higher the spectral order, the greater the overlap into the next order. The grating equation shows that the angles of the diffracted orders only depend on the grooves' period, and not on their shape. By controlling the cross-sectional profile of the grooves, it is possible to concentrate most of the diffracted energy in a particular order for a given wavelength. A triangular profile is commonly used. This technique is called blazing. The incident angle and wavelength for which the diffraction is most efficient are often called blazing angle and blazing wavelength. The efficiency of a grating may also depend on the polarization of the incident light. Gratings are usually designated by their groove density, the number of grooves per unit length, usually expressed in grooves per millimeter (g/mm), also equal to the inverse of the groove period. The groove period must be on the order of the wavelength of interest; the spectral range covered by a grating is dependent on groove spacing and is the same for ruled and holographic gratings with the same grating constant. The maximum wavelength that a grating can diffract is equal to twice the grating period, in which case the incident and diffracted light will be at ninety degrees to the grating normal. To obtain frequency dispersion over a wider frequency one must use a prism. In the optical regime, in which the use of gratings is most common, this corresponds to wavelengths between 100 nm and 10 µm. In that case, the groove density can vary from a few tens of grooves per millimeter, as in echelle gratings, to a few thousands of grooves per millimeter. When groove spacing is less than half the wavelength of light, the only present order is the m = 0 order. Gratings with such small periodicity are called subwavelength gratings and exhibit special optical properties. Made on an isotropic material the subwavelength gratings give rise to form birefringence, in which the material behaves as if it were birefringent. [edit] Fabrication Originally, high-resolution gratings were ruled using high-quality ruling engines whose construction was a large undertaking. Henry Joseph Grayson designed a machine to make diffraction gratings, succeeding with one of 120,000 lines to the inch (approx. 47 000 per cm) in Later, photolithographic techniques allowed gratings to be created from a holographic interference pattern. Holographic gratings have sinusoidal grooves and may not be as efficient as ruled gratings, but are often preferred in monochromators because they lead to much less stray light. A copying technique allows high quality replicas to be made from master gratings of either type, thereby lowering fabrication costs. Another method for manufacturing diffraction gratings uses a photosensitive gel sandwiched between two substrates. A holographic interference pattern exposes the gel which is later developed. These gratings, called volume phase holography diffraction gratings (or VPH diffraction gratings) have no physical grooves, bu

47 instead a periodic modulation of the refractive index within the gel
instead a periodic modulation of the refractive index within the gel. This removes much of the surface scattering effects typically seen in other types of gratings. These gratings also tend to have higher efficiencies, and allow for the inclusion of complicated patterns into a single grating. In older versions of such gratings, environmental susceptibility was a trade-off, as the gel had to be contained at low temperature and humidity. Typically, the photosensitive substances are sealed between two substrates which make them resistant to humidity, thermal and mechanical stresses. VPH diffraction gratings are not destroyed by accidental touches and are more scratch resistant than typical relief gratings. Semiconductor technology today is also utilized to etch holographically patterned gratings into robust materials such as fused silica. In this way, low stray-light holography is combined with the high efficiency of deep, etched transmission gratings, and can be incorporated into high volume, low cost semiconductor manufacturing technology. A new technology for grating insertion into integrated photonic lightwave circuits is digital planar holography (DPH). DPH gratings are generated in computer and fabricated on one or several interfaces of an optical waveguide planar with standard micro-lithography or nano-imprinting methods, compatible with mass-production. Light propagates inside the DPH gratings, confined by the refractive index gradient, which provides longer interaction path and greater flexibility in light steering. [edit] Examples The grooves of a compact disc can act as a grating and produce iridescent reflections. Diffraction gratings are often used in monochromators, spectrometers, lasers, wavelength division multiplexing devices, optical pulse compressing devices, and many other optical instruments. Ordinary pressed CD and DVD media are every-day examples of diffraction gratings and can be used to demonstrate the effect by reflecting sunlight off them onto a white wall. This is a side effect of their manufacture, as one surface of a CD has many small pits in the plastic, arranged in a spiral; that surface has a thin layer of metal applied to make the pits more visible. The structure of a DVD is optically similar, although it may have more than one pitted surface, and all pitted surfaces are inside the disc. In a standard pressed vinyl record when viewed from a low angle perpendicular to the grooves, a similar but less defined effect to that in a CD/DVD is seen. This is due to

48 [edit] See also [edit] References [edit] Natural gratings
viewing angle (less than the critical angle of reflection of the black vinyl) and the path of the light being reflected due to this being changed by the grooves, leaving a rainbow relief pattern behind. [edit] Natural gratings Striated muscle is the most commonly found natural diffraction grating[2] and, indeed, this has helped physiologists in determining the structure of such muscle. Aside from this, diffraction gratings are rarely present in nature. Most commonly confused with diffraction gratings are the iridescent colors of peacock feathers, mother-of-pearl, and butterfly wings. Iridescence is common in birds, fishes, insects, and some flowers, and is almost always caused by thin-film interference rather than diffraction.[3] Diffraction will produce the entire spectrum of colors as the viewing angle changes, whereas thin-film interference usually produces a much narrower range.[4] The cell structures in plants and animals are usually too irregular to produce the fine slit geometry necessary for a diffraction grating.[5] However, natural gratings do occur in some invertebrate marine animals, like the antennae of seed shrimp, and have even been discovered in Burgess Shale fossils.[6][7] Diffraction grating effects are sometimes seen in meteorology. Diffraction coronas are colorful rings surrounding a source of light, such as the sun. These are usually observed much closer to the light source than halos, and are caused by very fine particles, like water droplets, ice crystals, or smoke particles in a hazy sky. When the particles are all nearly the same size they diffract the incoming light at very specific angles. The exact angle depends on the size of the particles. Diffraction coronas are commonly observed around light sources, like candle flames or street lights, in the fog. Cloud iridescence is caused by diffraction, occurring along coronal rings when the particles in the clouds are all uniform in size.[8] [edit] See also Henry Augustus Rowland N-slit interferometric equation Zone plate [edit] References ^ Feynman, Richard (1985). QED: The Strange Theory of Light and Matter. Princeton, New Jersey: Princeton University Press. ^ Baskin et al. LIGHT DIFFRACTION STUDY OF SINGLE SKELETAL MUSCLE FIBERS. BIoPHYs.J.. PMC  ^ Nature's palette: the science of plant color By David Webster Lee - University of Chicago Press 2007 Page ^ Nature's palette: the science of plant color By David Webster Lee - University of Chicago Press 2007 Page 255 ^ Nature's palette: the science of plant color By David Webster Lee - University of Chicago Press 2007 Page 84 ^ Nature's palette: the science of plant color By David Webster Lee - University of Chicago Press 2007 Page

49 ^ http://wwwnhm.ac.uk/about-us/news/2006/mar/news_7834.html
^ Poarized light in nature By G. P. Können - Cambridge University Press 1985 Page 72-73  This article incorporates public domain material from the General Services Administration document "Federal Standard 1037C". Hutley, Michael, Diffraction Gratings (Techniques of Physics), Academic Press (1982). ISBN Loewen, Erwin & Evgeny Popov, Diffraction Gratings and Applications, CRC; 1 edition (1997). ISBN Palmer, Christopher, Diffraction Grating Handbook, 6th edition, Newport Corporation (2005). Greenslade, Thomas B., "Wire Diffraction Gratings," The Physics Teacher, February Volume 42 Issue 2, pp. 76–77. Abrahams, Peter, Early Instruments of Astronomical Spectroscopy. William E. L. Grossman, "The optical characteristics and production of diffraction gratings," Journal of Chemical Education 70:9 (Sep 1993), p. 741. National Optical Astronomy Observatories entry on volume phase holography gratings. [edit] External links Diffraction Gratings - The Crucial Dispersive Component Diffraction Grating Handbook Diffraction Grating Equations Automatic calculation of diffraction angles based on input variables. Optics Tutorial - Diffraction Gratings Ruled & Holographic Ray-Tracing program handling general reflective concave gratings for Windows XP and above View page ratings Rate this page What's this? Trustworthy Objective Complete Well-written I am highly knowledgeable about this topic (optional) Categories: Diffraction Optical devices Photonics Log in / create account Article Discussion

50 Diffractometer [edit] In principle
From Wikipedia, the free encyclopedia A diffractometer (pronunciation: di-"frak-'tä-m&-t&r) is a measuring instrument for analyzing the structure of a material from the scattering pattern produced when a beam of radiation or particles (such as X-rays or neutrons) interacts with it. [edit] In principle Because it is relatively easy to use electrons or neutrons having wavelengths smaller than a nanometer, electrons and neutrons may be used to study crystal structure in a manner very similar to X-ray diffraction. Electrons do not penetrate as deeply into matter as X-rays, hence electron diffraction reveals structure near the surface; neutrons do penetrate easily and have an advantage that they possess an intrinsic magnetic moment that causes them to interact differently with atoms having different alignments of their magnetic moments. This article does not cite any references or sources. Please help improve this article by adding citations to reliable sources. Unsourced material may be challenged and removed. (December 2009)

51 The detector end of a simple x-ray diffractometer with an area detector. The direction of the x rays is indicated with the red arrow. A typical diffractometer consists of a source of radiation, a monochromator to choose the wavelength, slits to adjust the shape of the beam, a sample and a detector. In a more complicated apparatus also a goniometer can be used for fine adjustment of the sample and the detector positions. When an area detector is used to monitor the diffracted radiation a beamstop is usually needed to stop the intense primary beam that has not been diffracted by the sample. Otherwise the detector might be damaged. Usually the beamstop can be completely impenetrable to the x rays or it may be semitransparent. The use of semitransparent beamstop allows the possibility to determine how much the sample absorbs the radiation using the intensity observed through the beamstop.


Download ppt "FURKAN KAMIL MOHAMMAD."

Similar presentations


Ads by Google