Fraunhofer diffraction

Last updated

In optics, the Fraunhofer diffraction equation is used to model the diffraction of waves when plane waves are incident on a diffracting object, and the diffraction pattern is viewed at a sufficiently long distance (a distance satisfying Fraunhofer condition) from the object (in the far-field region), and also when it is viewed at the focal plane of an imaging lens. [1] [2] In contrast, the diffraction pattern created near the diffracting object and (in the near field region) is given by the Fresnel diffraction equation.

Contents

The equation was named in honor of Joseph von Fraunhofer [3] although he was not actually involved in the development of the theory.[ citation needed ]

This article explains where the Fraunhofer equation can be applied, and shows Fraunhofer diffraction patterns for various apertures. A detailed mathematical treatment of Fraunhofer diffraction is given in Fraunhofer diffraction equation.

Equation

Example of far field (Fraunhofer) diffraction for a few aperture shapes.

When a beam of light is partly blocked by an obstacle, some of the light is scattered around the object, light and dark bands are often seen at the edge of the shadow – this effect is known as diffraction. [4] These effects can be modelled using the Huygens–Fresnel principle; Huygens postulated that every point on a wavefront acts as a source of spherical secondary wavelets and the sum of these secondary wavelets determines the form of the proceeding wave at any subsequent time, while Fresnel developed an equation using the Huygens wavelets together with the principle of superposition of waves, which models these diffraction effects quite well.

It is generally not straightforward to calculate the wave amplitude given by the sum of the secondary wavelets (The wave sum is also a wave.), each of which has its own amplitude, phase, and oscillation direction (polarization), since this involves addition of many waves of varying amplitude, phase, and polarization. When two light waves as electromagnetic fields are added together (vector sum), the amplitude of the wave sum depends on the amplitudes, the phases, and even the polarizations of individual waves. On a certain direction where electromagnetic wave fields are projected (or considering a situation where two waves have the same polarization), two waves of equal (projected) amplitude which are in phase (same phase) give the amplitude of the resultant wave sum as double the individual wave amplitudes, while two waves of equal amplitude which are in opposite phases give the zero amplitude of the resultant wave as they cancel out each other. Generally, a two-dimensional integral over complex variables has to be solved and in many cases, an analytic solution is not available. [5]

The Fraunhofer diffraction equation is a simplified version of Kirchhoff's diffraction formula and it can be used to model light diffraction when both a light source and a viewing plane (a plane of observation where the diffracted wave is observed) are effectively infinitely distant from a diffracting aperture. [6] With a sufficiently distant light source from a diffracting aperture, the incident light to the aperture is effectively a plane wave so that the phase of the light at each point on the aperture is the same. At a sufficiently distant plane of observation from the aperture, the phase of the wave coming from each point on the aperture varies linearly with the point position on the aperture, making the calculation of the sum of the waves at an observation point on the plane of observation relatively straightforward in many cases. Even the amplitudes of the secondary waves coming from the aperture at the observation point can be treated as same or constant for a simple diffraction wave calculation in this case. Diffraction in such a geometrical requirement is called Fraunhofer diffraction, and the condition where Fraunhofer diffraction is valid is called Fraunhofer condition, as shown in the right box. [7] A diffracted wave is often called Far field if it at least partially satisfies Fraunhofer condition such that the distance between the aperture and the observation plane is .

Fraunhofer diffraction occurs when:

(Fraunhofer condition)

– The largest size of a diffracting aperture or slit, – Wavelength, – The smaller of the two distances, one is between the diffracting aperture and the plane of observation and the other is between the diffracting plane and the point wave source.

For example, if a 0.5 mm diameter circular hole is illuminated by a laser light with 0.6 μm wavelength, then Fraunhofer diffraction occurs if the viewing distance is greater than 1000 mm.

Derivation of Fraunhofer condition

A geometrical diagram used to derive Fraunhofer condition at which Fraunhofer diffraction is valid. Fraunhofer Condition Derivation Geometry3.png
A geometrical diagram used to derive Fraunhofer condition at which Fraunhofer diffraction is valid.

The derivation of Fraunhofer condition here is based on the geometry described in the right box. [8] The diffracted wave path r2 can be expressed in terms of another diffracted wave path r1 and the distance b between two diffracting points by using the law of cosines;

This can be expanded by calculating the expression's Taylor series to second order with respect to ,

The phase difference between waves propagating along the paths r2 and r1 are, with the wavenumber where λ is the light wavelength,

If so , then the phase difference is . The geometrical implication from this expression is that the paths r2 and r1 are approximately parallel with each other. Since there can be a diffraction plane - observation plane diffracted wave path which angle with respect to a straight line parallel to the optical axis is close to 0, this approximation condition can be further simplified as where L is the distance between two planes along the optical axis. Due to the fact that an incident wave on a diffracting plane is effectively a plane wave if where L is the distance between the diffracting plane and the point wave source is satisfied, Fraunhofer condition is where L is the smaller of the two distances, one is between the diffracting plane and the plane of observation and the other is between the diffracting plane and the point wave source.

Focal plane of a positive lens as the far field plane

Plane wave focused by a lens. Lens and wavefronts rotated.gif
Plane wave focused by a lens.

In the far field, propagation paths for wavelets from every point on an aperture to a point of observation are approximately parallel, and a positive lens (focusing lens) focuses parallel rays toward the lens to a point on the focal plane (the focus point position on the focal plane depends on the angle of the parallel rays with respect to the optical axis). So, if a positive lens with a sufficiently long focal length (so that differences between electric field orientations for wavelets can be ignored at the focus) is placed after an aperture, then the lens practically makes the Fraunhofer diffraction pattern of the aperture on its focal plane as the parallel rays meet each other at the focus. [9]

Examples

In each of these examples, the aperture is illuminated by a monochromatic plane wave at normal incidence.

Diffraction by a narrow rectangular slit

Graph and image of single-slit diffraction Single Slit Diffraction (english).svg
Graph and image of single-slit diffraction

The width of the slit is W. The Fraunhofer diffraction pattern is shown in the image together with a plot of the intensity vs. angle θ. [10] The pattern has maximum intensity at θ = 0, and a series of peaks of decreasing intensity. Most of the diffracted light falls between the first minima. The angle, α, subtended by these two minima is given by: [11]

Thus, the smaller the aperture, the larger the angle α subtended by the diffraction bands. The size of the central band at a distance z is given by

Diffraction glass with 300 lines per millimeter Difraction glass.jpg
Diffraction glass with 300 lines per millimeter

For example, when a slit of width 0.5 mm is illuminated by light of wavelength 0.6 μm, and viewed at a distance of 1000 mm, the width of the central band in the diffraction pattern is 2.4 mm.

The fringes extend to infinity in the y direction since the slit and illumination also extend to infinity.

If W < λ, the intensity of the diffracted light does not fall to zero, and if D << λ, the diffracted wave is cylindrical.

Semi-quantitative analysis of single-slit diffraction

Geometry of single-slit diffraction Single slit diagram.svg
Geometry of single-slit diffraction

We can find the angle at which a first minimum is obtained in the diffracted light by the following reasoning. Consider the light diffracted at an angle θ where the distance CD is equal to the wavelength of the illuminating light. The width of the slit is the distance AC. The component of the wavelet emitted from the point A which is travelling in the θ direction is in anti-phase with the wave from the point B at middle of the slit, so that the net contribution at the angle θ from these two waves is zero. The same applies to the points just below A and B, and so on. Therefore, the amplitude of the total wave travelling in the direction θ is zero. We have:

The angle subtended by the first minima on either side of the centre is then, as above:

There is no such simple argument to enable us to find the maxima of the diffraction pattern.

Single-slit diffraction using Huygens' principle

Continuous broadside array of point sources of length a. Singleslithuygens.jpg
Continuous broadside array of point sources of length a.

We can develop an expression for the far field of a continuous array of point sources of uniform amplitude and of the same phase. Let the array of length a be parallel to the y axis with its center at the origin as indicated in the figure to the right. Then the differential field is: [12]

where . However and integrating from to ,

where .

Integrating we then get

Letting where the array length in radians is , then, [12]

Diffraction by a rectangular aperture

Computer simulation of Fraunhofer diffraction by a rectangular aperture Rectangular diffraction.jpg
Computer simulation of Fraunhofer diffraction by a rectangular aperture

The form of the diffraction pattern given by a rectangular aperture is shown in the figure on the right (or above, in tablet format). [13] There is a central semi-rectangular peak, with a series of horizontal and vertical fringes. The dimensions of the central band are related to the dimensions of the slit by the same relationship as for a single slit so that the larger dimension in the diffracted image corresponds to the smaller dimension in the slit. The spacing of the fringes is also inversely proportional to the slit dimension.

If the illuminating beam does not illuminate the whole vertical length of the slit, the spacing of the vertical fringes is determined by the dimensions of the illuminating beam. Close examination of the double-slit diffraction pattern below shows that there are very fine horizontal diffraction fringes above and below the main spot, as well as the more obvious horizontal fringes.

Diffraction by a circular aperture

Computer simulation of the Airy diffraction pattern Airy-pattern2.jpg
Computer simulation of the Airy diffraction pattern

The diffraction pattern given by a circular aperture is shown in the figure on the right. [14] This is known as the Airy diffraction pattern. It can be seen that most of the light is in the central disk. The angle subtended by this disk, known as the Airy disk, is

where W is the diameter of the aperture.

The Airy disk can be an important parameter in limiting the ability of an imaging system to resolve closely located objects.

Diffraction by an aperture with a Gaussian profile

Intensity of a plane wave diffracted through an aperture with a Gaussian profile Exp squared function.svg
Intensity of a plane wave diffracted through an aperture with a Gaussian profile

The diffraction pattern obtained given by an aperture with a Gaussian profile, for example, a photographic slide whose transmissivity has a Gaussian variation is also a Gaussian function. The form of the function is plotted on the right (above, for a tablet), and it can be seen that, unlike the diffraction patterns produced by rectangular or circular apertures, it has no secondary rings. [15] This technique can be used in a process called apodization—the aperture is covered by a Gaussian filter, giving a diffraction pattern with no secondary rings.

The output profile of a single mode laser beam may have a Gaussian intensity profile and the diffraction equation can be used to show that it maintains that profile however far away it propagates from the source. [16]

Diffraction by a double slit

Double-slit fringes with sodium light illumination SodiumD two double slits 2.jpg
Double-slit fringes with sodium light illumination

In the double-slit experiment, the two slits are illuminated by a single light beam. If the width of the slits is small enough (less than the wavelength of the light), the slits diffract the light into cylindrical waves. These two cylindrical wavefronts are superimposed, and the amplitude, and therefore the intensity, at any point in the combined wavefronts depends on both the magnitude and the phase of the two wavefronts. [17] These fringes are often known as Young's fringes.

The angular spacing of the fringes is given by

The spacing of the fringes at a distance z from the slits is given by [18]

where d is the separation of the slits.

The fringes in the picture were obtained using the yellow light from a sodium light (wavelength = 589 nm), with slits separated by 0.25 mm, and projected directly onto the image plane of a digital camera.

Double-slit interference fringes can be observed by cutting two slits in a piece of card, illuminating with a laser pointer, and observing the diffracted light at a distance of 1 m. If the slit separation is 0.5 mm, and the wavelength of the laser is 600 nm, then the spacing of the fringes viewed at a distance of 1 m would be 1.2 mm.

Semi-quantitative explanation of double-slit fringes

Geometry for far-field fringes Double slit.svg
Geometry for far-field fringes

The difference in phase between the two waves is determined by the difference in the distance travelled by the two waves.

If the viewing distance is large compared with the separation of the slits (the far field), the phase difference can be found using the geometry shown in the figure. The path difference between two waves travelling at an angle θ is given by

When the two waves are in phase, i.e. the path difference is equal to an integral number of wavelengths, the summed amplitude, and therefore the summed intensity is maximal, and when they are in anti-phase, i.e. the path difference is equal to half a wavelength, one and a half wavelengths, etc., then the two waves cancel, and the summed intensity is zero. This effect is known as interference.

The interference fringe maxima occur at angles

where λ is the wavelength of the light. The angular spacing of the fringes is given by

When the distance between the slits and the viewing plane is z, the spacing of the fringes is equal to and is the same as above:

Diffraction by a grating

Diffraction of a laser beam by a grating Diffraction of laser beam by grating.jpeg
Diffraction of a laser beam by a grating

A grating is defined in Born and Wolf as "any arrangement which imposes on an incident wave a periodic variation of amplitude or phase, or both".

A grating whose elements are separated by S diffracts a normally incident beam of light into a set of beams, at angles θn given by: [19]

This is known as the grating equation. The finer the grating spacing, the greater the angular separation of the diffracted beams.

If the light is incident at an angle θ0, the grating equation is:

The detailed structure of the repeating pattern determines the form of the individual diffracted beams, as well as their relative intensity while the grating spacing always determines the angles of the diffracted beams.

The image on the right shows a laser beam diffracted by a grating into n = 0, and ±1 beams. The angles of the first order beams are about 20°; if we assume the wavelength of the laser beam is 600 nm, we can infer that the grating spacing is about 1.8 μm.

Semi-quantitative explanation

Beugungsgitter.svg

A simple grating consists of a series of slits in a screen. If the light travelling at an angle θ from each slit has a path difference of one wavelength with respect to the adjacent slit, all these waves will add together, so that the maximum intensity of the diffracted light is obtained when:

This is the same relationship that is given above.

See also

Related Research Articles

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

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

<span class="mw-page-title-main">Wave interference</span> Phenomenon resulting from the superposition of two waves

In physics, interference is a phenomenon in which two coherent waves are combined by adding their intensities or displacements with due consideration for their phase difference. The resultant wave may have greater intensity or lower amplitude if the two waves are in phase or out of phase, respectively. Interference effects can be observed with all types of waves, for example, light, radio, acoustic, surface water waves, gravity waves, or matter waves as well as in loudspeakers as electrical waves.

<span class="mw-page-title-main">Wavelength</span> Distance over which a waves shape repeats

In physics and mathematics, wavelength or spatial period of a wave or periodic function is the distance over which the wave's shape repeats. In other words, it is the distance between consecutive corresponding points of the same phase on the wave, such as two adjacent crests, troughs, or zero crossings. Wavelength is a characteristic of both traveling waves and standing waves, as well as other spatial wave patterns. The inverse of the wavelength is called the spatial frequency. Wavelength is commonly designated by the Greek letter lambda (λ). The term "wavelength" is also sometimes applied to modulated waves, and to the sinusoidal envelopes of modulated waves or waves formed by interference of several sinusoids.

<span class="mw-page-title-main">Diffraction grating</span> Optical component which splits light into several beams

In optics, a diffraction grating is an optical grating with a periodic structure that diffracts light into several beams traveling in different directions. The emerging coloration is a form of structural coloration. The directions or diffraction angles of these beams depend on the wave (light) incident angle to the diffraction grating, the spacing or distance between adjacent diffracting elements on the grating, and the wavelength of the incident light. The grating acts as a dispersive element. Because of this, diffraction gratings are commonly used in monochromators and spectrometers, but other applications are also possible such as optical encoders for high-precision motion control and wavefront measurement.

<span class="mw-page-title-main">Snell's law</span> Formula for refraction angles

Snell's law is a formula used to describe the relationship between the angles of incidence and refraction, when referring to light or other waves passing through a boundary between two different isotropic media, such as water, glass, or air. In optics, the law is used in ray tracing to compute the angles of incidence or refraction, and in experimental optics to find the refractive index of a material. The law is also satisfied in meta-materials, which allow light to be bent "backward" at a negative angle of refraction with a negative refractive index.

<span class="mw-page-title-main">Arago spot</span> Bright point that appears at the center of a circular objects shadow due to Fresnel diffraction.

In optics, the Arago spot, Poisson spot, or Fresnel spot is a bright point that appears at the center of a circular object's shadow due to Fresnel diffraction. This spot played an important role in the discovery of the wave nature of light and is a common way to demonstrate that light behaves as a wave.

<span class="mw-page-title-main">Angular resolution</span> Ability of any image-forming device to distinguish small details of an object

Angular resolution describes the ability of any image-forming device such as an optical or radio telescope, a microscope, a camera, or an eye, to distinguish small details of an object, thereby making it a major determinant of image resolution. It is used in optics applied to light waves, in antenna theory applied to radio waves, and in acoustics applied to sound waves. The colloquial use of the term "resolution" sometimes causes confusion; when an optical system is said to have a high resolution or high angular resolution, it means that the perceived distance, or actual angular distance, between resolved neighboring objects is small. The value that quantifies this property, θ, which is given by the Rayleigh criterion, is low for a system with a high resolution. The closely related term spatial resolution refers to the precision of a measurement with respect to space, which is directly connected to angular resolution in imaging instruments. The Rayleigh criterion shows that the minimum angular spread that can be resolved by an image forming system is limited by diffraction to the ratio of the wavelength of the waves to the aperture width. For this reason, high resolution imaging systems such as astronomical telescopes, long distance telephoto camera lenses and radio telescopes have large apertures.

Fourier optics is the study of classical optics using Fourier transforms (FTs), in which the waveform being considered is regarded as made up of a combination, or superposition, of plane waves. It has some parallels to the Huygens–Fresnel principle, in which the wavefront is regarded as being made up of a combination of spherical wavefronts whose sum is the wavefront being studied. A key difference is that Fourier optics considers the plane waves to be natural modes of the propagation medium, as opposed to Huygens–Fresnel, where the spherical waves originate in the physical medium.

In many areas of science, Bragg's law, Wulff–Bragg's condition or Laue–Bragg interference, are a special case of Laue diffraction, giving the angles for coherent scattering of waves from a large crystal lattice. It describes how the superposition of wave fronts scattered by lattice planes leads to a strict relation between the wavelength and scattering angle. This law was initially formulated for X-rays, but it also applies to all types of matter waves including neutron and electron waves if there are a large number of atoms, as well as visible light with artificial periodic microscale lattices.

<span class="mw-page-title-main">Airy disk</span> Diffraction pattern in optics

In optics, the Airy disk and Airy pattern are descriptions of the best-focused spot of light that a perfect lens with a circular aperture can make, limited by the diffraction of light. The Airy disk is of importance in physics, optics, and astronomy.

<span class="mw-page-title-main">Acousto-optic modulator</span> Device which diffracts light via sound waves

An acousto-optic modulator (AOM), also called a Bragg cell or an acousto-optic deflector (AOD), uses the acousto-optic effect to diffract and shift the frequency of light using sound waves. They are used in lasers for Q-switching, telecommunications for signal modulation, and in spectroscopy for frequency control. A piezoelectric transducer is attached to a material such as glass. An oscillating electric signal drives the transducer to vibrate, which creates sound waves in the material. These can be thought of as moving periodic planes of expansion and compression that change the index of refraction. Incoming light scatters off the resulting periodic index modulation and interference occurs similar to Bragg diffraction. The interaction can be thought of as a three-wave mixing process resulting in sum-frequency generation or difference-frequency generation between phonons and photons.

<span class="mw-page-title-main">Fresnel diffraction</span> Diffraction

In optics, the Fresnel diffraction equation for near-field diffraction is an approximation of the Kirchhoff–Fresnel diffraction that can be applied to the propagation of waves in the near field. It is used to calculate the diffraction pattern created by waves passing through an aperture or around an object, when viewed from relatively close to the object. In contrast the diffraction pattern in the far field region is given by the Fraunhofer diffraction equation.

<span class="mw-page-title-main">Directivity</span> Measure of how much of an antennas signal is transmitted in one direction

In electromagnetics, directivity is a parameter of an antenna or optical system which measures the degree to which the radiation emitted is concentrated in a single direction. It is the ratio of the radiation intensity in a given direction from the antenna to the radiation intensity averaged over all directions. Therefore, the directivity of a hypothetical isotropic radiator is 1, or 0 dBi.

<span class="mw-page-title-main">Acousto-optics</span> The study of sound and light interaction

Acousto-optics is a branch of physics that studies the interactions between sound waves and light waves, especially the diffraction of laser light by ultrasound through an ultrasonic grating.

<span class="mw-page-title-main">Diffraction from slits</span>

Diffraction processes affecting waves are amenable to quantitative description and analysis. Such treatments are applied to a wave passing through one or more slits whose width is specified as a proportion of the wavelength. Numerical approximations may be used, including the Fresnel and Fraunhofer approximations.

<span class="mw-page-title-main">Contrast transfer function</span>

The contrast transfer function (CTF) mathematically describes how aberrations in a transmission electron microscope (TEM) modify the image of a sample. This contrast transfer function (CTF) sets the resolution of high-resolution transmission electron microscopy (HRTEM), also known as phase contrast TEM.

Quantum mechanics was first applied to optics, and interference in particular, by Paul Dirac. Richard Feynman, in his Lectures on Physics, uses Dirac's notation to describe thought experiments on double-slit interference of electrons. Feynman's approach was extended to N-slit interferometers for either single-photon illumination, or narrow-linewidth laser illumination, that is, illumination by indistinguishable photons, by Frank Duarte. The N-slit interferometer was first applied in the generation and measurement of complex interference patterns.

<span class="mw-page-title-main">Thin-film interference</span> Optical phenomenon

Thin-film interference is a natural phenomenon in which light waves reflected by the upper and lower boundaries of a thin film interfere with one another, increasing reflection at some wavelengths and decreasing it at others. When white light is incident on a thin film, this effect produces colorful reflections.

In optics, the Fraunhofer diffraction equation is used to model the diffraction of waves when the diffraction pattern is viewed at a long distance from the diffracting object, and also when it is viewed at the focal plane of an imaging lens.

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

References

  1. Born & Wolf 1999 , p. 427
  2. Jenkins & White 1957 , p. 288
  3. "Fraunhofer, Joseph von (1787-1826) -- from Eric Weisstein's World of Scientific Biography".
  4. Heavens, O. S.; Ditchburn, R. W. (1991). Insight into Optics. Chichester: Longman and Sons. p. 62. ISBN   0-471-92769-4. OCLC   22114471.
  5. Born & Wolf 1999 , p. 425
  6. Jenkins & White 1957 , Section 15.1, p. 288
  7. Lipson, A.; Lipson, S. G.; Lipson, H. (2011). Optical physics (4th ed.). Cambridge: Cambridge University Press. p. 203. ISBN   978-0-521-49345-1. OCLC   637708967.
  8. Hecht, Eugene (2017). "Problem 9.21". Optics (5th ed.). Pearson. p. 453. ISBN   978-1-292-09693-3.
  9. Hecht 2002 , p. 448
  10. Hecht 2002 , Figures 10.6(b) and 10.7(e)
  11. Jenkins & White 1957 , p. 297
  12. 1 2 Kraus, John Daniel; Marhefka, Ronald J. (2002). Antennas for all applications. McGraw-Hill. ISBN   9780072321036.
  13. Born & Wolf 1999 , Figure 8.10
  14. Born & Wolf 1999 , Figure 8.12
  15. Hecht 2002 , Figure 11.33
  16. Hecht 2002 , Figure 13.14
  17. Born & Wolf 1999 , Figure 7.4
  18. Hecht 2002 , eq. (9.30).
  19. Longhurst, R. S. (1967). Geometrical and Physical Optics (2nd ed.). London: Longmans. eq.(12.1).

Sources