User:Jordgette/doubleslit1

From Wikipedia, the free encyclopedia

The double-slit experiment, sometimes called Young's experiment, is a demonstration that matter and energy can display characteristics of both waves and particles. In the basic version of the experiment, a coherent light source such as a laser beam illuminates a thin plate with two parallel slits, and the light passing through the slits is observed on a screen behind the plate. The wave nature of light causes the light waves passing through the two slits to interfere, producing bright and dark bands on the screen — a result that would not be expected if light consisted strictly of particles. However, at the screen, the light is always found to be absorbed as though it were composed of discrete particles or photons.[1][2] This establishes the principle known as wave–particle duality.

Overview[edit]

Same double-slit assembly (0.7mm between slits); in top image, one slit is closed. Note that the single-slit diffraction pattern — the faint spots on either side of the main band — is also seen in the double-slit image, but at twice the intensity and with the addition of many smaller interference fringes.

If light consisted strictly of ordinary or classical particles, and these particles were fired in a straight line through a slit and allowed to strike a screen on the other side, we would expect to see a pattern corresponding to the size and shape of the slit. However, when this "single-slit experiment" is actually performed, the pattern on the screen is a diffraction pattern, a fairly narrow central band with dimmer bands parallel to it on each side. (See the top photograph to the right.)

Similarly, if light consisted strictly of classical particles and we illuminated two parallel slits, the expected pattern on the screen would simply be the sum of the two single-slit patterns. In actuality, however, the pattern becomes wider and much more detailed, including a series of light and dark bands. (See the bottom photograph to the right.) When Thomas Young first demonstrated this phenomenon, it indicated that light consists of waves, as the distribution of brightness can be explained by the alternately additive and subtractive interference of wavefronts.[3] Young's experiment played a vital part in the acceptance of the wave theory of light in the early 1800s, vanquishing the corpuscular theory of light proposed by Isaac Newton, which had been the accepted model of light propagation in the 17th and 18th centuries. However, the later discovery of the photoelectric effect demonstrated that under different circumstances, light can behave as if it is composed of discrete particles. These seemingly contradictory discoveries made it necessary to go beyond classical physics and take the quantum nature of light into account.

A double-slit experiment was not performed with anything other than light until 1961, when Clauss Jönsson of the University of Tübingen performed it with electrons.[4][5] In 2002, the double-slit experiment of Claus Jönsson was voted "the most beautiful experiment" by readers of Physics World.[6]

In 1999, objects large enough to be seen under a microscopebuckyball molecules (diameter about 0.7 nm, nearly half a million times larger than a proton) — were found to exhibit wave-like interference.[7][8]

The double-slit experiment (and its variations) has become a classic thought experiment for its clarity in expressing the central puzzles of quantum mechanics. Richard Feynman was fond of saying that all of quantum mechanics can be gleaned from carefully thinking through the implications of this single experiment.[9]

Variations of the experiment[edit]

Interference of individual particles[edit]

Electron buildup over time

An important version of this experiment involves single particles (or waves — for consistency, they are called particles here). Sending particles through a double-slit apparatus one at a time results in single particles appearing on the screen, as expected. Remarkably, however, an interference pattern emerges when these particles are allowed to build up one by one (see the image to the right). For example, when a laboratory apparatus was developed that could reliably fire one electron at a time through the double slit,[10] the emergence of an interference pattern suggested that each electron was interfering with itself, and therefore in some sense the electron had to be going through both slits at once[11] — an idea that contradicts our everyday experience of discrete objects. This phenomenon has also been shown to occur with atoms and even some molecules, including buckyballs.[7][12][13]

When electrons (for example) are fired singly through a double-slit apparatus, they do not cluster around two single points corresponding to the two slits, but instead, one by one, they create an interference pattern. This experimental fact is highly reproducible, and the mathematics of quantum mechanics (see below) allows us to predict the exact probability of an electron striking the screen at any particular point. However, the electrons do not arrive at the screen in any predictable order. In other words, knowing where all the previous electrons appeared on the screen and in what order tells us nothing about where any future electron will hit, even though the probabilities at specific points can be calculated.[14] Thus, we have the appearance of a seemingly causeless selection event in a highly orderly and predictable formulation of the interference pattern. This has prompted some theorists to look for additional determinants or "hidden variables" in the system which, were they to become known, would account for the location of each individual impact with the target.[15]

A low intensity double-slit experiment was first performed by Taylor in 1909,[16] by reducing the level of incident light until on average only one photon was being transmitted at a time. The appearance of interference built up from individual photons could be explained by understanding that a single photon has its own wavefront that passes through both slits, and that the single photon will show up on the detector screen according to the net probability values resulting from the co-incidence of the two probability waves coming by way of the two slits.[17] Note that it is the probabilities of photons appearing at various points along the detection screen that add or cancel. So, if there is a cancellation of waves at some point, that does not mean that a photon disappears; it only means that the probability of a photon's appearing at that point will decrease, and the probability that it will appear somewhere else increases.

A detailed treatment of the mathematics of double-slit interference in the context of quantum mechanics is given in the article on Englert-Greenberger duality.

With particle detectors at the slits[edit]

The double-slit apparatus can be modified by adding particle detectors positioned at the slits. This enables the experimenter to find the position of a particle not when it impacts the screen, but rather, when it passes through the double-slit — did it go through only one of the slits, as a particle would be expected to do, or through both, as a wave would be expected to do? Numerous experiments have shown, however, that any modification of the apparatus that can determine which slit a particle passes through weakens the interference pattern at the screen,[3] illustrating the complementarity principle: that light (and electrons, etc.) can behave as either particles or waves, but not both at the same time.[18][19][20] An experiment performed in 1987[21] produced results that demonstrated that information could be obtained regarding which path a particle had taken, without destroying the interference altogether. This showed the effect of measurements that disturbed the particles in transit to a lesser degree and thereby influenced the interference pattern only to a comparable extent.

There are many methods to determine whether a photon passed through a slit, for instance by placing an atom at the position of each slit. Interesting experiments of this latter kind have been performed with photons[21] and with neutrons.[22]

Delayed choice & quantum eraser[edit]

The delayed-choice experiment and the quantum eraser are sophisticated variations of the double-slit with particle detectors placed not at the slits but elsewhere in the apparatus. The first demonstrates that extracting "which path" information after a particle passes through the slits can seem to retroactively alter its previous behavior at the slits. The second demonstrates that wave behavior can be restored by erasing or otherwise making permanently unavailable the "which path" information.

Other variations[edit]

A laboratory double-slit assembly; distance between top posts approximately one inch.

In 1926 Max Born proposed that as a consequence of the quantum mechanics, only two slits would produce the familiar results of the double-slit experiment, while three or more slits would not.[23] In July 2010 this prediction was tested by Sinha and colleagues and found to be correct.[24] Sinha and colleagues made three parallel slits, each 300 μm tall, and 30 μm wide, in a stainless steel plate.[25] They were able to independently open or close a shutter over each slit, giving rise to eight possible states.[25] The interference pattern obtained with all three slits open was subtracted from the six states with one slit or with two slits open, and "resulted in a number very close to zero,"[25] thus confirming the Born hypothesis.[23] (The eighth state is all three slits closed.)

In 1967 Pfleegor and Mandel demonstrated two-source interference using two separate lasers as light sources.[26][27]

It was shown experimentally in 1972 that in a double-slit system where only one slit was open at any time, interference was nonetheless observed provided the path difference was such that the detected photon could have come from either slit.[28] [29] The experimental conditions were such that the photon density in the system was much less than unity.

Classical wave optics & mathematical formulation[edit]

Huygens-Fresnel wave propagation

Christiaan Huygens understood the basic idea of how light propagates and how to predict its path through a physical apparatus. He understood that a light source emits a series of waves comparable to the way that water waves spread out from something like a fishing float that is jiggled up and down and bobs on the water surface. He said that the way to predict where the next wave front will be found is to generate a series of concentric circles on a sufficiently large number of points on a known wave front and then draw a curve that will pass tangent to all the resulting circles out in front of the known wave front. The diagram given here shows what happens when a flat wave front is extended in this manner, and what happens when a curved wave front is extended in the same way. Augustin Fresnel (1788–1827) based his proof that the wave nature of light does not contradict the observed fact that light propagates in a straight line in homogeneous media on Huygens' work, and also based himself on Huygens' ideas to give a complete account of diffraction and interference phenomena known at his time.[17] The diagram shows what happens when a flat wave front encounters a series of slits in a wall. Following the same principle elucidated above, it is clear that the new wave front will "bulge out" from each slit and light will be experienced as having diverged around the edges of the slit.

Two slits are illuminated by a plane wave

The second figure illustrates how two-slit interference occurs in terms of classical wave propagation. Two slits are illuminated by a plane wave. Light is diffracted by each of the slits. These two 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. The relative distance travelled by the two wavefronts varies as we move in the x-direction varies, so the relative phase also varies. As it changes from 0 to 2π, the magnitude of the summed waves fluctuates between minimum and maximum values. It is this fluctuation which is observed as interference fringes. The spacing of the fringes depends on the geometry of the arrangement.

An expression for the intensity of the diffracted light field can be calculated using the Fraunhofer diffraction equation. If the width of the slits is negligible, their separation is d, and they are illuminated normally by a plane wave with wavelength λ, the intensity variation with angle θ, which is the angle subtended by the point P at the origin, is given by

It can be seen that the intensity of the pattern varies as the square of the cosine, thus giving rise to Young's fringes. The spacing of the fringes increases as the separation of the slits decreases.

The bright bands observed on the screen happen when the light has interfered constructively—where a crest of a wave meets a crest from another wave. The dark regions show destructive interference—a crest meets a trough. Constructive interference occurs when

and destructive interference occurs when

Using the paraxial approximation, when θ < 10°, that , the bright fringes occur when

where

n is the order of maximum observed (central maximum is n = 0),
x is the distance between the bands of light and the central maximum (also called fringe distance),
L is the distance from the slits to the screen centerpoint, and
θn is the angle between the centerpoint normal and the nth maximum.

A more complete discussion can be found here [30]

It is possible to work out the wavelength of light using this equation and the above apparatus. If d and L are known and x is observed, then λ can be easily calculated.

If the width of the slits, a is finite, the equation for the diffracted pattern is given by Longhurst [31] as

Again the intensity varies as the square of the cosine, but it is now modulated by the diffraction pattern of the individual slits.

Similar expressions can be calculated for slits of finite depth, two point sources, two circular sources, etc. using the Fraunhofer diffraction equation. In each case, the intensity will vary as the square of the sine which is modulated by another function.

Similar calculations can be done using the Fresnel diffraction equation. As the plane of observation gets closer to the plane in which the slits are located, the diffraction patterns associated with each slit decrease in size, so that the interference between the two patterns is reduced, and may vanish altogether when there is no overlap in the two diffracted patterns.[31]

Interpretations of the experiment[edit]

Like the Schrödinger's cat thought experiment, the double-slit experiment is often used to highlight the differences and similarities between the various interpretations of quantum mechanics.

Copenhagen interpretation[edit]

The Copenhagen interpretation is a consensus among some of the pioneers in the field of quantum mechanics that it is undesirable to posit anything that goes beyond the mathematical formulae and the kinds of physical apparatus and reactions that enable us to gain some knowledge of what goes on at the atomic scale. One of the mathematical constructs that enables experimenters to predict very accurately certain experimental results is sometimes called a probability wave. In its mathematical form it is analogous to the description of a physical wave, but its "crests" and "troughs" indicate levels of probability for the occurrence of certain phenomena (e.g., a spark of light at a certain point on a detector screen) that can be observed in the macro world of ordinary human experience.

The probability "wave" can be said to "pass through space" because the probability values that one can compute from its mathematical representation are dependent on time. One cannot speak of the location of any particle such as a photon between the time it is emitted and the time it is detected simply because in order to say that something is located somewhere at a certain time one has to detect it (of course, since photons travel at a known speed (the speed of light) at any given time (stated to Planck accuracy) you can calculate (to within Planck distance) where the 'probability' field is 'centered', but until the particle is detected, you can not be certain 'exactly' where it is. The requirement for the eventual appearance of an interference pattern is that particles be emitted, and that there be a screen with at least two slits between the emitter and the detection screen. Experiments observe nothing whatsoever between the time of emission of the particle and its arrival at the detection screen. However, it is essential that both slits be an equal distance from the center line, and that they be within a certain maximum distance of each other that is related to the wavelength of the particle being emitted. If a ray tracing is then made as if a light wave (as understood in classical physics) is wide enough to encounter both slits and passes through both of them, then that ray tracing will accurately predict the appearance of maxima and minima on the detector screen when many particles pass through the apparatus and gradually "paint" the expected interference pattern.

Path-integral formulation[edit]

One of an infinite number of equally likely paths used in the Feynman path integral. (see also: Wiener process.)

The Copenhagen interpretation is similar to the path integral formulation of quantum mechanics provided by Feynman. The path integral formulation replaces the classical notion of a single, unique trajectory for a system, with a sum over all possible trajectories. The trajectories are added together by using functional integration.

Each path is considered equally likely, and thus contributes the same amount. However, the phase of this contribution at any given point along the path is determined by the action along the path (see Euler's formula):

All these contributions are then added together, and the magnitude of the final result is squared, to get the probability distribution for the position of a particle:

As is always the case when calculating probability, the results must then be normalized:

To summarize, the probability distribution of the outcome is the normalized square of the norm of the superposition, over all paths from the point of origin to the final point, of waves propagating proportionally to the action along each path. The differences in the cumulative action along the different paths (and thus the relative phases of the contributions) produces the interference pattern observed by the double-slit experiment. Feynman stressed that his formulation is merely a mathematical description, not an attempt to describe a real process that we cannot measure.

Relational interpretation[edit]

According to the relational interpretation of quantum mechanics, first proposed by Carlo Rovelli,[32] observations such as those in the double-slit experiment result specifically from the interaction between the observer and the object being observed, not any absolute property possessed by the object. In the case of an electron, if it is initially observed at a particular slit, then the observer/particle interaction includes information about the electron's position. This partially constrains the particle's eventual location at the screen. If it is observed not at a particular slit but rather at the screen, then there is no "which slit" information as part of the interaction, so the electron's observed position on the screen is determined strictly by its probability function. This makes the resulting pattern on the screen the same as if each individual electron had passed through both slits. It has also been suggested that space and distance themselves are relational, and that an electron can appear to be in "two places at once" — e.g., at both slits — because its spatial relations to particular points on the screen remain identical from both slit locations.[33]

See also[edit]

References[edit]

  1. ^ Feynman, Richard P. (1965). The Feynman Lectures on Physics, Vol. 3. USA: Addison-Wesley. pp. 1–8. ISBN 0201021188P. {{cite book}}: Check |isbn= value: invalid character (help)
  2. ^ Darling, David (2007). "Wave - Particle Duality". The Internet Encyclopedia of Science. The Worlds of David Darling. Retrieved 18 October 2008.
  3. ^ a b Feynman, Richard P. (1965). The Feynman Lectures on Physics. Massachusetts, USA: Addison-Wesley. pp. 1–1 to 1–9. ISBN 0201021188P. {{cite book}}: Check |isbn= value: invalid character (help); Unknown parameter |coauthors= ignored (|author= suggested) (help)
  4. ^ Jönsson C,(1961) Zeitschrift für Physik, 161:454–474
  5. ^ Jönsson C (1974). Electron diffraction at multiple slits. American Journal of Physics, 4:4–11.
  6. ^ "The most beautiful experiment". Physics World 2002.
  7. ^ a b New Scientist: Quantum wonders: Corpuscles and buckyballs, 2010 (Introduction, subscription needed for full text, quoted in full in [1])
  8. ^ Nature: Wave–particle duality of C60 molecules, 14 October 1999. Abstract, subscription needed for full text
  9. ^ Greene, Brian (1999). The Elegant Universe: Superstrings, Hidden Dimensions, and the Quest for the Ultimate Theory. New York: W.W. Norton. pp. 97–109. ISBN 0393046885.
  10. ^ O Donati G F Missiroli G Pozzi May 1973 An Experiment on Electron Interference American Journal of Physics 41 639–644
  11. ^ Brian Greene, The Elegant Universe, p. 110
  12. ^ http://prl.aps.org/abstract/PRL/v87/i16/e160401
  13. ^ Nairz O, Arndt M, and Zeilinger A. Quantum interference experiments with large molecules. American Journal of Physics, 2003; 71:319-325. http://scitation.aip.org/getabs/servlet/GetabsServlet?prog=normal&id=AJPIAS000071000004000319000001&idtype=cvips&gifs=yes
  14. ^ Brian Greene, The Elegant Universe, p. 104, pp. 109-114
  15. ^ Greene, Brian (2004). The Fabric of the Cosmos: Space, Time, and the Texture of Reality. Knopf. pp. 204–213. ISBN 0375412883.
  16. ^ Sir Geoffrey Ingram Taylor, "Interference Fringes with Feeble Light", Proc. Cam. phil. Soc. 15, 114 (1909).
  17. ^ a b de Broglie, Louis (1953). The revolution in physics; a non-mathematical survey of quanta. Translated by Ralph W. Niemeyer. New York: Noonday Press. pp. 47, 117, 178–186.
  18. ^ Harrison, David (2002). "Complementarity and the Copenhagen Interpretation of Quantum Mechanics". UPSCALE. Dept. of Physics, U. of Toronto. Retrieved 21 June 2008.
  19. ^ Cassidy, David (2008). "Quantum Mechanics 1925–1927: Triumph of the Copenhagen Interpretation". Werner Heisenberg. American Institute of Physics. Retrieved 21 June 2008.
  20. ^ Boscá Díaz-Pintado, María C. (29–31 March 2007). "Updating the wave-particle duality". 15th UK and European Meeting on the Foundations of Physics. Leeds, UK. Retrieved 21 June 2008. {{cite conference}}: Cite has empty unknown parameter: |coauthors= (help)
  21. ^ a b Mittelstaedt, P.; Prieur, A.; Schieder, R. (1987). "Unsharp particle-wave duality in a photon split-beam experiment". Foundations of Physics. 17 (9): 891–903. Bibcode:1987FoPh...17..891M. doi:10.1007/BF00734319.{{cite journal}}: CS1 maint: date and year (link) Also D.M. Greenberger and A. Yasin, "Simultaneous wave and particle knowledge in a neutron interferometer", Physics Letters A 128, 391-4 (1988).
  22. ^ Summhammer, J.; Rauch, H.; Tuppinger, D. (1987). "Stochastic and deterministic absorption in neutron-interference experiments". Phys. Rev. A. 36 (9): 4447–4455. Bibcode:1987PhRvA..36.4447S. doi:10.1103/PhysRevA.36.4447. PMID 9899403.{{cite journal}}: CS1 maint: date and year (link)
  23. ^ a b Born, Max (1926). "Zur Quantenmechanik der Stoßvorgänge". Zeitschrift für Physik A Hadrons and Nuclei. 37 (12): 863–867. Bibcode:1926ZPhy...37..863B. doi:10.1007/BF01397477.
  24. ^ Sinha, Urbasi; Couteau, Christophe; Jennewein, Thomas; Laflamme, Raymond; Weihs, G (23 July 2010). "Ruling Out Multi-Order Interference in Quantum Mechanics". Science. 329 (5990): 418–421. arXiv:1007.4193. Bibcode:2010Sci...329..418S. doi:10.1126/science.1190545. PMID 20651147. {{cite journal}}: Unknown parameter |unused_data= ignored (help)
  25. ^ a b c Sanders, Laura; Couteau, C; Jennewein, T; Laflamme, R; Weihs, G (14 August 2010). "Triple slits don't add interference". Science News. 178 (4): 418–421. arXiv:1007.4193. Bibcode:2010Sci...329..418S. doi:10.1126/science.1190545. PMID 20651147.
  26. ^ Pfleegor, R. L. and Mandel, L. (July 1967). "Interference of Independent Photon Beams". Phys. Rev. 159 (5): 1084–1088. doi:10.1103/PhysRev.159.1084.{{cite journal}}: CS1 maint: date and year (link) CS1 maint: multiple names: authors list (link)
  27. ^ http://scienceblogs.com/principles/2010/11/interference_of_independent_ph.php>
  28. ^ Sillitto, R.M. and Wykes, Catherine (1972). "An interference experiment with light beams modulated in anti-phase by an electro-optic shutter". Physics Letters A. 39 (4): 333–334. doi:10.1016/0375-9601(72)91015-8.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  29. ^ http://www.sillittopages.co.uk/80rms_35.html "To a light particle"
  30. ^ For a more complete discussion, with diagrams and photographs, see Arnold L Reimann, Physics, chapter 38.
  31. ^ a b RS Longhurst, Geometrical and Physical Optics, 1967, Longmans, London
  32. ^ Rovelli, Carlo (1996). "Relational Quantum Mechanics". International Journal of Theoretical Physics. 35 (8): 1637–1678. arXiv:quant-ph/9609002. Bibcode:1996IJTP...35.1637R. doi:10.1007/BF02302261.
  33. ^ Filk, Thomas (2006). "Relational Interpretation of the Wave Function and a Possible Way Around Bell's Theorem". International Journal of Theoretical Physics. 45 (6): 1205–1219. doi:10.1007/s10773-006-9125-0.

Further reading[edit]

  • Feynman, Richard P. (1988). QED: The Strange Theory of Light and Matter. Princeton University Press. ISBN 0-691-02417-0.
  • Frank, Philipp (1957). Philosophy of Science. Prentice-Hall.
  • Gribbin, John (1999). Q is for Quantum: Particle Physics from A to Z. Weidenfeld & Nicolson. ISBN 0-7538-0685-1.
  • French, A.P. and Edwin F. Taylor (1978). An Introduction to Quantum Physics. Norton. ISBN 0-393-09106-6.
  • Greene, Brian (2000). The Elegant Universe. Vintage. ISBN 0-375-70811-1.
  • Greene, Brian (2005). The Fabric of the Cosmos. Vintage. ISBN 0-375-72720-5.
  • Hey, Tony (2003). The New Quantum Universe. Cambridge University Press. ISBN 0-5215-6457-3.
  • Sears, Francis Weston (1949). Optics. Addison Wesley.
  • Tipler, Paul (2004). Physics for Scientists and Engineers: Electricity, Magnetism, Light, and Elementary Modern Physics (5th ed.). W. H. Freeman. ISBN 0-7167-0810-8.
  • Al-Khalili, Jim (2003). Quantum: A Guide for the Perplexed. Weidenfeld and Nicholson, London. ISBN 0-297-84305-2.

External links[edit]


[[Category:Foundational quantum physics]] [[Category:Physics experiments]] [[Category:Quantum mechanics]] [[Category:Wave mechanics]]