In physics, the hydrodynamic quantum analogs refer to experimentally-observed phenomena involving bouncing fluid droplets over a vibrating fluid bath that behave analogously to several quantum-mechanical systems. The experimental evidence for diffraction through slits has been disputed, however, though the diffraction pattern of walking droplets is not
exactly the same as in quantum physics, it does appear clearly in the
high memory parameter regime (at high forcing of the bath) where all the
quantum-like effects are strongest.
A droplet can be made to bounce indefinitely in a stationary
position on a vibrating fluid surface. This is possible due to a
pervading air layer that prevents the drop from coalescing into the bath. For certain combinations of bath surface acceleration, droplet size, and vibration frequency, a bouncing droplet will cease to stay in a stationary position, but instead “walk” in a rectilinear motion on top of the fluid bath. Walking droplet systems have been found to mimic several quantum mechanical phenomena including particle diffraction, quantum tunneling, quantized orbits, the Zeeman Effect, and the quantum corral.
Besides being an interesting means to visualise phenomena that
are typical of the quantum-mechanical world, floating droplets on a
vibrating bath have interesting analogies with the pilot wave theory,
one of the many interpretations of quantum mechanics in its early
stages of conception and development. The theory was initially proposed
by Louis de Broglie in 1927.
It suggests that all particles in motion are actually borne on a
wave-like motion, similar to how an object moves on a tide. In this
theory, it is the evolution of the carrier wave that is given by the Schrödinger equation. It is a deterministic theory and is entirely nonlocal. It is an example of a hidden variable theory,
and all non-relativistic quantum mechanics can be accounted for in this
theory. The theory was abandoned by de Broglie in 1932, gave way to the
Copenhagen interpretation, but was revived by David Bohm in 1952 as De Broglie–Bohm theory.
The Copenhagen interpretation does not use the concept of the carrier
wave or that a particle moves in definite paths until a measurement is
made.
Physics of bouncing and walking droplets
History
Floating droplets on a vibrating bath were first described in writing by Jearl Walker in a 1978 article in Scientific American.
In 2005, Yves Couder and his lab were the first to systematically
study the dynamics of bouncing droplets and discovered most of the
quantum mechanical analogs.
John Bush and his lab expanded upon Couder's work and studied the
system in greater detail. In 2015 three separate groups, including John
Bush, attempted to reproduce the effect and were unsuccessful.
Stationary bouncing droplet
A
fluid droplet can float or bounce over a vibrating fluid bath because
of the presence of an air layer between the droplet and the bath
surface. The behavior of the droplet depends on the acceleration
of the bath surface. Below a critical acceleration, the droplet will
take successively smaller bounces before the intervening air layer
eventually drains from underneath, causing the droplet to coalesce.
Above the bouncing threshold, the intervening air layer replenishes
during each bounce so the droplet never touches the bath surface. Near
the bath surface, the droplet experiences equilibrium
between inertial forces, gravity, and a reaction force due to the
interaction with the air layer above the bath surface. This reaction
force serves to launch the droplet back above the air like a trampoline. Molacek and Bush proposed two different models for the reaction force.
Walking droplet
For a small range of frequencies
and drop sizes, a fluid droplet on a vibrating bath can be made to
“walk” on the surface if the surface acceleration is sufficiently high
(but still below the Faraday
instability). That is, the droplet does not simply bounce in a
stationary position but instead wanders in a straight line or in a
chaotic trajectory. When a droplet interacts with the surface, it
creates a transient wave that propagates from the point of impact. These
waves usually decay, and stabilizing forces keep the droplet from
drifting. However, when the surface acceleration is high, the transient
waves created upon impact do not decay as quickly, deforming the surface
such that the stabilizing forces are not enough to keep the droplet
stationary. Thus, the droplet begins to “walk.”
Quantum phenomena on a macroscopic scale
A walking droplet on a vibrating fluid bath was found to behave
analogously to several different quantum mechanical systems, namely
particle diffraction, quantum tunneling, quantized orbits, the Zeeman effect, and the quantum corral.
Single and double slit diffraction
It has been known since the early 19th century that when light is shone through one or two small slits, a diffraction
pattern appears on a screen far from the slits. Light has wave-like
behavior, and interferes with itself through the slits, creating a
pattern of alternating high and low intensity. Single electrons also exhibit wave-like behavior as a result of wave-particle duality. When electrons are fired through small slits, the probability of the electron striking the screen at a specific point shows an interference pattern as well.
In 2006, Couder and Fort demonstrated that walking droplets
passing through one or two slits exhibit similar interference behavior.
They used a square shaped vibrating fluid bath with a constant depth
(aside from the walls). The “walls” were regions of much lower depth,
where the droplets would be stopped or reflected away. When the droplets
were placed in the same initial location, they would pass through the
slits and be scattered, seemingly randomly. However, by plotting a histogram
of the droplets based on scattering angle, the researchers found that
the scattering angle was not random, but droplets had preferred
directions that followed the same pattern as light or electrons. In this
way, the droplet may mimic the behavior of a quantum particle as it passes through the slit.
Despite that research, in 2015 three teams: Bohr and Andersen's group in Denmark, Bush's team at MIT, and a team led by the quantum physicist Herman Batelaan at the University of Nebraska
set out to repeat the Couder and Fort's bouncing-droplet double-slit
experiment. Having their experimental setups perfected, none of the
teams saw the interference-like pattern reported by Couder and Fort. Droplets went through the slits in almost straight lines, and no stripes appeared.
It has since been shown that droplet trajectories are sensitive
to interactions with container boundaries, air currents, and other
parameters. Though the diffraction pattern of walking droplets is not
exactly the same as in quantum physics, and is not expected to show a
Fraunhofer-like dependence of the number of peaks on the slit width, the
diffraction pattern does appear clearly in the high memory regime (at
high forcing of the bath).
Quantum tunneling
Quantum tunneling
is the quantum mechanical phenomenon where a quantum particle passes
through a potential barrier. In classical mechanics, a classical
particle could not pass through a potential barrier if the particle does
not have enough energy, so the tunneling effect is confined to the
quantum realm. For example, a rolling ball would not reach the top of a
steep hill without adequate energy. However, a quantum particle, acting
as a wave, can undergo both reflection and transmission at a potential
barrier. This can be shown as a solution to the time dependent Schrödinger Equation.
There is a finite, but usually small, probability to find the electron
at a location past the barrier. This probability decreases exponentially
with increasing barrier width.
The macroscopic analogy using fluid droplets was first
demonstrated in 2009. Researchers set up a square vibrating bath
surrounded by walls on its perimeter. These “walls” were regions of
lower depth, where a walking droplet may be reflected away. When the
walking droplets were allowed to move around in the domain, they usually
were reflected away from the barriers. However, surprisingly, sometimes
the walking droplet would bounce past the barrier, similar to a quantum
particle undergoing tunneling. In fact, the crossing probability was
also found to decrease exponentially with increasing width of the
barrier, exactly analogous to a quantum tunneling particle.
Quantized orbits
When two atomic particles interact and form a bound state, such the hydrogen atom,
the energy spectrum is discrete. That is, the energy levels of the
bound state are not continuous and only exist in discrete quantities,
forming “quantized orbits.” In the case of a hydrogen atom, the
quantized orbits are characterized by atomic orbitals, whose shapes are functions of discrete quantum numbers.
On the macroscopic level, two walking fluid droplets can interact
on a vibrating surface. It was found that the droplets would orbit each
other in a stable configuration with a fixed distance apart. The stable
distances came in discrete values. The stable orbiting droplets
analogously represent a bound state in the quantum mechanical system.
The discrete values of the distance between droplets are analogous to
discrete energy levels as well.
Zeeman effect
When an external magnetic field
is applied to a hydrogen atom, for example, the energy levels are
shifted to values slightly above or below the original level. The
direction of shift depends on the sign of the z-component of the total
angular momentum. This phenomenon is known as the Zeeman Effect.
In the context of walking droplets, an analogous Zeeman Effect
can be demonstrated by observing orbiting droplets in a vibrating fluid
bath.
The bath is also brought to rotate at a constant angular velocity. In
the rotating bath, the equilibrium distance between droplets shifts
slightly farther or closer. The direction of shift depends on whether
the orbiting drops rotate in the same direction as the bath or in
opposite directions. The analogy to the quantum effect is clear. The
bath rotation is analogous to an externally applied magnetic field, and
the distance between droplets is analogous to energy levels. The
distance shifts under an applied bath rotation, just as the energy
levels shift under an applied magnetic field.
Quantum corral
Researchers
have found that a walking droplet placed in a circular bath does not
wander randomly, but rather there are specific locations the droplet is
more likely to be found. Specifically, the probability of finding the
walking droplet as a function of the distance from the center is
non-uniform and there are several peaks of higher probability. This probability distribution mimics that of an electron confined to a quantum corral.
Photons or matter (like electrons) produce an interference pattern when two slits are used
Light from a green laser passing through two slits 0.4 mm wide and 0.1 mm apart
In modern physics, the double-slit experiment demonstrates that light and matter can exhibit behavior of both classical particles and classical waves. This type of experiment was first performed by Thomas Young in 1801, as a demonstration of the wave behavior of visible light. In 1927, Davisson and Germer and, independently, George Paget Thomson and his research student Alexander Reid demonstrated that electrons show the same behavior, which was later extended to atoms and molecules. Thomas Young's experiment with light was part of classical physics long before the development of quantum mechanics and the concept of wave–particle duality. He believed it demonstrated that the Christiaan Huygens'wave theory of light was correct, and his experiment is sometimes referred to as Young's experiment or Young's slits.
The experiment belongs to a general class of "double path" experiments,
in which a wave is split into two separate waves (the wave is typically
made of many photons and better referred to as a wave front, not to be
confused with the wave properties of the individual photon) that later
combine into a single wave. Changes in the path-lengths of both waves
result in a phase shift, creating an interference pattern. Another version is the Mach–Zehnder interferometer, which splits the beam with a beam splitter.
In the basic version of this experiment, a coherent light source, such as a laser
beam, illuminates a plate pierced by 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 of classical particles.
However, the light is always found to be absorbed at the screen at
discrete points, as individual particles (not waves); the interference
pattern appears via the varying density of these particle hits on the
screen. Furthermore, versions of the experiment that include detectors at the slits find that each detected photon passes through one slit (as would a classical particle), and not through both slits (as would a wave). However, such experiments
demonstrate that particles do not form the interference pattern if one
detects which slit they pass through. These results demonstrate the
principle of wave–particle duality.
Other atomic-scale entities, such as electrons, are found to exhibit the same behavior when fired towards a double slit.
Additionally, the detection of individual discrete impacts is observed
to be inherently probabilistic, which is inexplicable using classical mechanics.
The experiment can be done with entities much larger than
electrons and photons, although it becomes more difficult as size
increases. The largest entities for which the double-slit experiment has
been performed were molecules that each comprised 2000 atoms (whose total mass was 25,000 atomic mass units).
The double-slit experiment (and its variations) has become a
classic for its clarity in expressing the central puzzles of quantum
mechanics. Richard Feynman called it "a phenomenon which is impossible […] to explain in any classical way, and which has in it the heart of quantum mechanics. In reality, it contains the only mystery [of quantum mechanics]."
Overview
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
in which the light is spread out. The smaller the slit, the greater the
angle of spread. The top portion of the image shows the central portion
of the pattern formed when a red laser illuminates a slit and, if one
looks carefully, two faint side bands. More bands can be seen with a
more highly refined apparatus. Diffraction explains the pattern as being the result of the interference of light waves from the slit.
If one illuminates two parallel slits, the light from the two
slits again interferes. Here the interference is a more pronounced
pattern with a series of alternating light and dark bands. The width of
the bands is a property of the frequency of the illuminating light. (See the bottom photograph to the right.)
When Thomas Young
(1773–1829) 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.
Young's experiment, performed in the early 1800s, played a crucial role
in the understanding of the wave theory of light, 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
into account the quantum nature of light.
Feynman was fond of saying that all of quantum mechanics can be
gleaned from carefully thinking through the implications of this single
experiment.
He also proposed (as a thought experiment) that if detectors were
placed before each slit, the interference pattern would disappear.
The Englert–Greenberger duality relation provides a detailed treatment of the mathematics of double-slit interference in the context of quantum mechanics.
A low-intensity double-slit experiment was first performed by G. I. Taylor in 1909, by reducing the level of incident light until photon emission/absorption events were mostly non-overlapping.
A slit interference experiment was not performed with anything other than light until 1961, when Claus Jönsson of the University of Tübingen performed it with coherent electron beams and multiple slits. In 1974, the Italian physicists Pier Giorgio Merli, Gian Franco Missiroli, and Giulio Pozzi
performed a related experiment using single electrons from a coherent
source and a biprism beam splitter, showing the statistical nature of
the buildup of the interference pattern, as predicted by quantum theory. In 2002, the single-electron version of the experiment was voted "the most beautiful experiment" by readers of Physics World. Since that time a number of related experiments have been published, with a little controversy.
In 2012, Stefano Frabboni and co-workers sent single electrons
onto nanofabricated slits (about 100 nm wide) and, by detecting the
transmitted electrons with a single-electron detector, they could show
the build-up of a double-slit interference pattern.
Many related experiments involving the coherent interference have been
performed; they are the basis of modern electron diffraction, microscopy
and high resolution imaging.
In 2018, single particle interference was demonstrated for antimatter in the Positron Laboratory (L-NESS, Politecnico di Milano) of Rafael Ferragut in Como (Italy), by a group led by Marco Giammarchi.
Variations of the experiment
Interference from individual particles
An
important version of this experiment involves single particle
detection. Illuminating the double-slit with a low intensity results in
single particles being detected as white dots on the screen.
Remarkably, however, an interference pattern emerges when these
particles are allowed to build up one by one (see the image below).
Experimental electron double slit diffraction pattern.
Across the middle of the image at the top, the intensity alternates
from high to low, showing interference in the signal from the two slits.
Bottom: movie of the pattern being built up dot-by-dot.
This demonstrates the wave–particle duality,
which states that all matter exhibits both wave and particle
properties: The particle is measured as a single pulse at a single
position, while the modulus squared of the wave describes the probability of detecting the particle at a specific place on the screen giving a statistical interference pattern. This phenomenon has been shown to occur with photons, electrons, atoms, and even some molecules: with buckminsterfullerene (C 60) in 2001, with 2 molecules of 430 atoms (C 60(C 12F 25) 10 and C 168H 94F 152O 8N 4S 4) in 2011, and with molecules of up to 2000 atoms in 2019.
In addition to interference patterns built up from single particles, up to 4 entangled photons can also show interference patterns.
The Mach–Zehnder interferometer can be seen as a simplified version of the double-slit experiment.
Instead of propagating through free space after the two slits, and
hitting any position in an extended screen, in the interferometer the
photons can only propagate via two paths, and hit two discrete
photodetectors. This makes it possible to describe it via simple linear
algebra in dimension 2, rather than differential equations.
A photon emitted by the laser hits the first beam splitter and is
then in a superposition between the two possible paths. In the second
beam splitter these paths interfere, causing the photon to hit the
photodetector on the right with probability one, and the photodetector
on the bottom with probability zero.
Blocking one of the paths, or equivalently detecting the presence of a
photon on a path eliminates interference between the paths: both
photodetectors will be hit with probability 1/2. This indicates that
after the first beam splitter the photon does not take one path or
another, but rather exists in a quantum superposition of the two paths.
"Which-way" experiments and the principle of complementarity
A well-known thought experiment
predicts that if particle detectors are positioned at the slits,
showing through which slit a photon goes, the interference pattern will
disappear. This which-way experiment illustrates the complementarity principle that photons can behave as either particles or waves, but cannot be observed as both at the same time.
Despite the importance of this thought experiment in the history of quantum mechanics (for example, see the discussion on Einstein's version of this experiment), technically feasible realizations of this experiment were not proposed until the 1970s.
(Naive implementations of the textbook thought experiment are not
possible because photons cannot be detected without absorbing the
photon.) Currently, multiple experiments have been performed
illustrating various aspects of complementarity.
An experiment performed in 1987
produced results that demonstrated that partial information could be
obtained regarding which path a particle had taken without destroying
the interference altogether. This "wave-particle trade-off" takes the
form of an inequality relating the visibility of the interference pattern and the distinguishability of the which-way paths.
Wheeler's delayed-choice experiments
demonstrate that extracting "which path" information after a particle
passes through the slits can seem to retroactively alter its previous
behavior at the slits.
Quantum eraser
experiments demonstrate that wave behavior can be restored by erasing
or otherwise making permanently unavailable the "which path"
information.
A simple do-it-at-home illustration of the quantum eraser phenomenon was given in an article in Scientific American.
If one sets polarizers before each slit with their axes orthogonal to
each other, the interference pattern will be eliminated. The polarizers
can be considered as introducing which-path information to each beam.
Introducing a third polarizer in front of the detector with an axis of
45° relative to the other polarizers "erases" this information, allowing
the interference pattern to reappear. This can also be accounted for by
considering the light to be a classical wave, and also when using circular polarizers and single photons. Implementations of the polarizers using entangled photon pairs have no classical explanation.
In a highly publicized experiment in 2012, researchers claimed to
have identified the path each particle had taken without any adverse
effects at all on the interference pattern generated by the particles.
In order to do this, they used a setup such that particles coming to
the screen were not from a point-like source, but from a source with two
intensity maxima. However, commentators such as Svensson have pointed out that there is in fact no conflict between the weak measurements performed in this variant of the double-slit experiment and the Heisenberg uncertainty principle.
Weak measurement followed by post-selection did not allow simultaneous
position and momentum measurements for each individual particle, but
rather allowed measurement of the average trajectory of the particles
that arrived at different positions. In other words, the experimenters
were creating a statistical map of the full trajectory landscape.
Other variations
In 1967, Pfleegor and Mandel demonstrated two-source interference using two separate lasers as light sources.
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. The experimental conditions were such that the photon density in the system was much less than 1.
In 1991, Carnal and Mlynek performed the classic Young's double slit experiment with metastable helium atoms passing through micrometer-scale slits in gold foil.
In 1999, a quantum interference experiment (using a diffraction
grating, rather than two slits) was successfully performed with
buckyball molecules (each of which comprises 60 carbon atoms). A buckyball is large enough (diameter about 0.7 nm, nearly half a million times larger than a proton) to be seen in an electron microscope.
In 2002, an electron field emission source was used to
demonstrate the double-slit experiment. In this experiment, a coherent
electron wave was emitted from two closely located emission sites on the
needle apex, which acted as double slits, splitting the wave into two
coherent electron waves in a vacuum. The interference pattern between
the two electron waves could then be observed.
In 2017, researchers performed the double-slit experiment using
light-induced field electron emitters. With this technique, emission
sites can be optically selected on a scale of ten nanometers. By
selectively deactivating (closing) one of the two emissions (slits),
researchers were able to show that the interference pattern disappeared.
In 2005, E. R. Eliel presented an experimental and theoretical
study of the optical transmission of a thin metal screen perforated by
two subwavelength slits, separated by many optical wavelengths. The
total intensity of the far-field double-slit pattern is shown to be
reduced or enhanced as a function of the wavelength of the incident
light beam.
In 2012, researchers at the University of Nebraska–Lincoln performed the double-slit experiment with electrons as described by Richard Feynman,
using new instruments that allowed control of the transmission of the
two slits and the monitoring of single-electron detection events.
Electrons were fired by an electron gun and passed through one or two
slits of 62 nm wide × 4 μm tall.
In 2013, a quantum interference experiment (using diffraction
gratings, rather than two slits) was successfully performed with
molecules that each comprised 810 atoms (whose total mass was over
10,000 atomic mass units). The record was raised to 2000 atoms (25,000 amu) in 2019.
Hydrodynamic pilot wave analogs
Hydrodynamic analogs
have been developed that can recreate various aspects of quantum
mechanical systems, including single-particle interference through a
double-slit.
A silicone oil droplet, bouncing along the surface of a liquid,
self-propels via resonant interactions with its own wave field. The
droplet gently sloshes the liquid with every bounce. At the same time,
ripples from past bounces affect its course. The droplet's interaction
with its own ripples, which form what is known as a pilot wave,
causes it to exhibit behaviors previously thought to be peculiar to
elementary particles – including behaviors customarily taken as evidence
that elementary particles are spread through space like waves, without
any specific location, until they are measured.
Behaviors mimicked via this hydrodynamic pilot-wave system include quantum single particle diffraction,
tunneling, quantized orbits, orbital level splitting, spin, and
multimodal statistics. It is also possible to infer uncertainty
relations and exclusion principles. Videos are available illustrating
various features of this system.
However, more complicated systems that involve two or more
particles in superposition are not amenable to such a simple,
classically intuitive explanation. Accordingly, no hydrodynamic analog of entanglement has been developed. Nevertheless, optical analogs are possible.
Double-slit experiment on time
In 2023, an experiment was reported recreating an interference pattern in time by shining a pump laser pulse at a screen coated in indium tin oxide (ITO) which would alter the properties of the electrons within the material due to the Kerr effect,
changing it from transparent to reflective for around 200 femtoseconds
long where a subsequent probe laser beam hitting the ITO screen would
then see this temporary change in optical properties as a slit in time
and two of them as a double slit with a phase difference adding up
destructively or constructively on each frequency component resulting in
an interference pattern. Similar results have been obtained classically on water waves.
Classical wave-optics formulation
Much of the behaviour of light can be modelled using classical wave theory. The Huygens–Fresnel principle
is one such model; it states that each point on a wavefront generates a
secondary wavelet, and that the disturbance at any subsequent point can
be found by summing the contributions of the individual wavelets at that point. This summation needs to take into account the phase as well as the amplitude of the individual wavelets. Only the intensity of a light field can be measured—this is proportional to the square of the amplitude.
In the double-slit experiment, the two slits are illuminated by
the quasi-monochromatic light of a single laser. If the width of the
slits is small enough (much less than the wavelength of the laser
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. 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 below right. The path difference between two waves travelling at
an angle θ is given by:
Where d is the distance between the two slits. 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
maximum, 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, θf, is given by
The spacing of the fringes at a distance z from the slits is given by
For example, if two slits are separated by 0.5 mm (d), and are illuminated with a 0.6 μm wavelength laser (λ), then at a distance of 1 m (z), the spacing of the fringes will be 1.2 mm.
If the width of the slits b is appreciable compared to the wavelength, the Fraunhofer diffraction equation is needed to determine the intensity of the diffracted light as follows:
where the sinc function is defined as sinc(x) = sin(x)/x for x ≠ 0, and sinc(0) = 1.
This is illustrated in the figure above, where the first pattern is the diffraction pattern of a single slit, given by the sinc
function in this equation, and the second figure shows the combined
intensity of the light diffracted from the two slits, where the cos
function represents the fine structure, and the coarser structure
represents diffraction by the individual slits as described by the sinc function.
Similar calculations for the near field can be made by applying the Fresnel diffraction
equation, which implies that 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 area in which
interference occurs is reduced, and may vanish altogether when there is
no overlap in the two diffracted patterns.
Path-integral formulation
The double-slit experiment can illustrate 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:
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 by imposing:
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 wavespropagatingproportionally
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 can measure.
The
standard interpretation of the double slit experiment is that the
pattern is a wave phenomenon, representing interference between two
probability amplitudes, one for each slit. Low intensity experiments
demonstrate that the pattern is filled in one particle detection at a
time. Any change to the apparatus designed to detect a particle at a
particular slit alters the probability amplitudes and the interference
disappears. This interpretation is independent of any conscious observer.
Niels Bohr interpreted quantum experiments like the double-slit experiment using the concept of complementarity.
In Bohr's view quantum systems are not classical, but measurements can
only give classical results. Certain pairs of classical properties will
never be observed in a quantum system simultaneously: the interference
pattern of waves in the double slit experiment will disappear if
particles are detected at the slits. Modern quantitative versions of the
concept allow for a continuous tradeoff between the visibility of the
interference fringes and the probability of particle detection at a
slit.
The Copenhagen interpretation is a collection of views about the meaning of quantum mechanics, stemming from the work of Niels Bohr, Werner Heisenberg, Max Born,
and others. The term "Copenhagen interpretation" was apparently coined
by Heisenberg during the 1950s to refer to ideas developed in the
1925–1927 period, glossing over his disagreements with Bohr.Consequently, there is no definitive historical statement of what the
interpretation entails. Features common across versions of the
Copenhagen interpretation include the idea that quantum mechanics is
intrinsically indeterministic, with probabilities calculated using the Born rule, and some form of complementarity principle. Moreover, the act of "observing" or "measuring" an object is irreversible, and no truth can be attributed to an object, except according to the results of its measurement.
In the Copenhagen interpretation, complementarity means a particular
experiment can demonstrate particle behavior (passing through a definite
slit) or wave behavior (interference), but not both at the same time. In a Copenhagen-type view, the question of which slit a particle travels through has no meaning when there is no detector.
Relational interpretation
According to the relational interpretation of quantum mechanics, first proposed by Carlo Rovelli, observations such as those in the double-slit experiment result specifically from the interaction between the observer
(measuring device) and the object being observed (physically interacted
with), 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 (photon–electron) interaction includes
information about the electron's position. This partially constrains the
particle's eventual location at the screen. If it is "observed"
(measured with a photon) not at a particular slit but rather at the
screen, then there is no "which path" 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.
Many-worlds interpretation
As with Copenhagen, there are multiple variants of the many-worlds interpretation.
The unifying theme is that physical reality is identified with a
wavefunction, and this wavefunction always evolves unitarily, i.e.,
following the Schrödinger equation with no collapses. Consequently, there are many parallel universes, which only interact with each other through interference. David Deutsch
argues that the way to understand the double-slit experiment is that in
each universe the particle travels through a specific slit, but its
motion is affected by interference with particles in other universes,
and this interference creates the observable fringes.
David Wallace, another advocate of the many-worlds interpretation,
writes that in the familiar setup of the double-slit experiment the two
paths are not sufficiently separated for a description in terms of
parallel universes to make sense.
An alternative to the standard understanding of quantum mechanics, the De Broglie–Bohm theory
states that particles also have precise locations at all times, and
that their velocities are defined by the wave-function. So while a
single particle will travel through one particular slit in the
double-slit experiment, the so-called "pilot wave" that influences it
will travel through both. The two slit de Broglie-Bohm trajectories were
first calculated by Chris Dewdney while working with Chris Philippidis
and Basil Hiley at Birkbeck College (London).
The de Broglie-Bohm theory produces the same statistical results as
standard quantum mechanics, but dispenses with many of its conceptual
difficulties by adding complexity through an ad hoc quantum potential to guide the particles.
More complex variants of this type of approach have appeared, for instance the three wave hypothesis of Ryszard Horodecki as well as other complicated combinations of de Broglie and Compton waves. To date there is no evidence that these are useful.
Bohmian trajectories
Trajectories of particles in De Broglie–Bohm theory in the double-slit experiment.
100
trajectories guided by the wave function. In De Broglie-Bohm's theory, a
particle is represented, at any time, by a wave function and a position (center of mass). This is a kind of augmented reality compared to the standard interpretation.
Numerical
simulation of the double-slit experiment with electrons. Figure on the
left: evolution (from left to right) of the intensity of the electron
beam at the exit of the slits (left) up to the detection screen located
10 cm after the slits (right). The higher the intensity, the more the
color is light blue – Figure in the center: impacts of the electrons
observed on the screen – Figure on the right: intensity of the electrons
in the far field
approximation (on the screen). Numerical data from Claus Jönsson's
experiment (1961). Photons, atoms and molecules follow a similar
evolution.
Coherence expresses the potential for two waves to interfere. Two monochromatic beams from a single source always interfere. Wave sources are not strictly monochromatic: they may be partly coherent. Beams from different sources are mutually incoherent.
When interfering, two waves add together to create a wave of greater amplitude than either one (constructive interference) or subtract from each other to create a wave of minima which may be zero (destructive interference), depending on their relative phase.
Constructive or destructive interference are limit cases, and two
waves always interfere, even if the result of the addition is
complicated or not remarkable.
Two waves with constant relative phase will be coherent. The amount of coherence can readily be measured by the interference visibility,
which looks at the size of the interference fringes relative to the
input waves (as the phase offset is varied); a precise mathematical
definition of the degree of coherence
is given by means of correlation functions. More broadly, coherence
describes the statistical similarity of a field, such as an
electromagnetic field or quantum wave packet, at different points in
space or time.
Qualitative concept
Coherence controls the visibility or contrast of interference patterns. For example, visibility of the double slit experiment pattern requires that both slits be illuminated by a coherent wave as illustrated in the figure. Large sources without collimation or sources that mix many different frequencies will have lower visibility.
Coherence contains several distinct concepts. Spatial coherence
describes the correlation (or predictable relationship) between waves
at different points in space, either lateral or longitudinal. Temporal coherence describes the correlation between waves observed at different moments in time. Both are observed in the Michelson–Morley experiment and Young's interference experiment. Once the fringes are obtained in the Michelson interferometer,
when one of the mirrors is moved away gradually from the beam-splitter,
the time for the beam to travel increases and the fringes become dull
and finally disappear, showing temporal coherence. Similarly, in a double-slit experiment,
if the space between the two slits is increased, the coherence dies
gradually and finally the fringes disappear, showing spatial coherence.
In both cases, the fringe amplitude slowly disappears, as the path
difference increases past the coherence length.
The coherence function between two signals and is defined as
where is the cross-spectral density of the signal and and are the power spectral density functions of and , respectively. The cross-spectral density and the power spectral density are defined as the Fourier transforms of the cross-correlation and the autocorrelation
signals, respectively. For instance, if the signals are functions of
time, the cross-correlation is a measure of the similarity of the two
signals as a function of the time lag relative to each other and the
autocorrelation is a measure of the similarity of each signal with
itself in different instants of time. In this case the coherence is a
function of frequency. Analogously, if and
are functions of space, the cross-correlation measures the similarity
of two signals in different points in space and the autocorrelations the
similarity of the signal relative to itself for a certain separation
distance. In that case, coherence is a function of wavenumber (spatial frequency).
The coherence varies in the interval . If it means that the signals are perfectly correlated or linearly related and if they are totally uncorrelated. If a linear system is being measured, being the input and
the output, the coherence function will be unitary all over the
spectrum. However, if non-linearities are present in the system the
coherence will vary in the limit given above.
Coherence and correlation
The coherence of two waves expresses how well correlated the waves are as quantified by the cross-correlation function.
Cross-correlation quantifies the ability to predict the phase of the
second wave by knowing the phase of the first. As an example, consider
two waves perfectly correlated for all times (by using a monochromatic
light source). At any time, the phase difference between the two waves
will be constant. If, when they are combined, they exhibit perfect
constructive interference, perfect destructive interference, or
something in-between but with constant phase difference, then it follows
that they are perfectly coherent. As will be discussed below, the
second wave need not be a separate entity. It could be the first wave at
a different time or position. In this case, the measure of correlation
is the autocorrelation function (sometimes called self-coherence). Degree of correlation involves correlation functions.
Examples of wave-like states
These states are unified by the fact that their behavior is described by a wave equation or some generalization thereof.
Waves in a rope (up and down) or slinky (compression and expansion)
In system with macroscopic waves, one can measure the wave directly.
Consequently, its correlation with another wave can simply be
calculated. However, in optics one cannot measure the electric field directly as it oscillates much faster than any detector's time resolution. Instead, one measures the intensity
of the light. Most of the concepts involving coherence which will be
introduced below were developed in the field of optics and then used in
other fields. Therefore, many of the standard measurements of coherence
are indirect measurements, even in fields where the wave can be measured
directly.
Temporal coherence
Temporal coherence is the measure of the average correlation between the value of a wave and itself delayed by ,
at any pair of times. Temporal coherence tells us how monochromatic a
source is. In other words, it characterizes how well a wave can
interfere with itself at a different time. The delay over which the
phase or amplitude wanders by a significant amount (and hence the
correlation decreases by significant amount) is defined as the coherence time. At a delay of the degree of coherence is perfect, whereas it drops significantly as the delay passes . The coherence length is defined as the distance the wave travels in time .
The coherence time is not the time duration of the signal; the coherence length differs from the coherence area (see below).
The relationship between coherence time and bandwidth
The larger the bandwidth – range of frequencies Δf a wave contains – the faster the wave decorrelates (and hence the smaller is):
Formally, this follows from the convolution theorem in mathematics, which relates the Fourier transform of the power spectrum (the intensity of each frequency) to its autocorrelation.
Narrow bandwidth lasers have long coherence lengths (up to hundreds of meters). For example, a stabilized and monomode helium–neon laser can easily produce light with coherence lengths of 300 m. Not all lasers have a high monochromaticity, however (e.g. for a mode-locked Ti-sapphire laser, Δλ ≈ 2 nm – 70 nm).
LEDs are characterized by Δλ ≈ 50 nm, and tungsten filament
lights exhibit Δλ ≈ 600 nm, so these sources have shorter coherence
times than the most monochromatic lasers.
Examples of temporal coherence
Examples of temporal coherence include:
A wave containing only a single frequency (monochromatic) is
perfectly correlated with itself at all time delays, in accordance with
the above relation. (See Figure 1)
Conversely, a wave whose phase drifts quickly will have a short coherence time. (See Figure 2)
Similarly, pulses (wave packets)
of waves, which naturally have a broad range of frequencies, also have a
short coherence time since the amplitude of the wave changes quickly.
(See Figure 3)
Finally, white light, which has a very broad range of frequencies,
is a wave which varies quickly in both amplitude and phase. Since it
consequently has a very short coherence time (just 10 periods or so), it
is often called incoherent.
Holography requires light with a long coherence time. In contrast, optical coherence tomography, in its classical version, uses light with a short coherence time.
Measurement of temporal coherence
In optics, temporal coherence is measured in an interferometer such as the Michelson interferometer or Mach–Zehnder interferometer. In these devices, a wave is combined with a copy of itself that is delayed by time . A detector measures the time-averaged intensity
of the light exiting the interferometer. The resulting visibility of
the interference pattern (e.g. see Figure 4) gives the temporal
coherence at delay .
Since for most natural light sources, the coherence time is much
shorter than the time resolution of any detector, the detector itself
does the time averaging. Consider the example shown in Figure 3. At a
fixed delay, here , an infinitely fast detector would measure an intensity that fluctuates significantly over a time t equal to . In this case, to find the temporal coherence at , one would manually time-average the intensity.
Spatial coherence
In
some systems, such as water waves or optics, wave-like states can
extend over one or two dimensions. Spatial coherence describes the
ability for two spatial points x1 and x2
in the extent of a wave to interfere when averaged over time. More
precisely, the spatial coherence is the cross-correlation between two
points in a wave for all times. If a wave has only 1 value of amplitude
over an infinite length, it is perfectly spatially coherent. The range
of separation between the two points over which there is significant
interference defines the diameter of the coherence area, (Coherence length ,
often a feature of a source, is usually an industrial term related to
the coherence time of the source, not the coherence area in the medium).
is the relevant type of coherence for the Young's double-slit
interferometer. It is also used in optical imaging systems and
particularly in various types of astronomy telescopes.
A distance away from an incoherent source with surface area ,
Sometimes people also use "spatial coherence" to refer to the
visibility when a wave-like state is combined with a spatially shifted
copy of itself.
Consider a tungsten light-bulb filament. Different points in the
filament emit light independently and have no fixed phase-relationship.
In detail, at any point in time the profile of the emitted light is
going to be distorted. The profile will change randomly over the
coherence time . Since for a white-light source such as a light-bulb is small, the filament is considered a spatially incoherent source. In contrast, a radio antenna array,
has large spatial coherence because antennas at opposite ends of the
array emit with a fixed phase-relationship. Light waves produced by a
laser often have high temporal and spatial coherence (though the degree
of coherence depends strongly on the exact properties of the laser).
Spatial coherence of laser beams also manifests itself as speckle
patterns and diffraction fringes seen at the edges of shadow.
Holography requires temporally and spatially coherent light. Its inventor, Dennis Gabor,
produced successful holograms more than ten years before lasers were
invented. To produce coherent light he passed the monochromatic light
from an emission line of a mercury-vapor lamp through a pinhole spatial filter.
In February 2011 it was reported that helium atoms, cooled to near absolute zero / Bose–Einstein condensate state, can be made to flow and behave as a coherent beam as occurs in a laser. Moreover, the coherence properties of the output light from multimode
nonlinear optical structures were found to obey the optical
thermodynamic theory.
Spectral coherence of short pulses
Waves of different frequencies (in light these are different colours)
can interfere to form a pulse if they have a fixed relative
phase-relationship (see Fourier transform).
Conversely, if waves of different frequencies are not coherent, then,
when combined, they create a wave that is continuous in time (e.g. white
light or white noise). The temporal duration of the pulse is limited by the spectral bandwidth of the light according to:
If the phase depends linearly on the frequency (i.e. ) then the pulse will have the minimum time duration for its bandwidth (a transform-limited pulse), otherwise it is chirped (see dispersion).
Light also has a polarization,
which is the direction in which the electric or magnetic field
oscillates. Unpolarized light is composed of incoherent light waves with
random polarization angles. The electric field of the unpolarized light
wanders in every direction and changes in phase over the coherence time
of the two light waves. An absorbing polarizer rotated to any angle will always transmit half the incident intensity when averaged over time.
If the electric field wanders by a smaller amount the light will
be partially polarized so that at some angle, the polarizer will
transmit more than half the intensity. If a wave is combined with an
orthogonally polarized copy of itself delayed by less than the coherence
time, partially polarized light is created.
The polarization of a light beam is represented by a vector in the Poincaré sphere.
For polarized light the end of the vector lies on the surface of the
sphere, whereas the vector has zero length for unpolarized light. The
vector for partially polarized light lies within the sphere.
The signature property of quantum matter waves,
wave interference, relies on coherence. While initially patterned after
optical coherence, the theory and experimental understanding of quantum
coherence greatly expanded the topic.
Matter wave coherence
The simplest extension of optical coherence applies optical concepts to matter waves. For example, when performing the double-slit experiment
with atoms instead of light waves, a sufficiently collimated atomic
beam creates a coherent atomic wave-function illuminating both slits.
Each slit acts as a separate but in-phase beam contributing to the
intensity pattern on a screen. These two contributions give rise to an
intensity pattern of bright bands due to constructive interference,
interlaced with dark bands due to destructive interference, on a
downstream screen. Many variations of this experiment have been
demonstrated.
As with light, transverse coherence (across the direction of
propagation) of matter waves is controlled by collimation. Because
light, at all frequencies, travels the same velocity, longitudinal and
temporal coherence are linked; in matter waves these are independent. In
matter waves, velocity (energy) selection controls longitudinal
coherence and pulsing or chopping controls temporal coherence.
Quantum optics
The discovery of the Hanbury Brown and Twiss effect – correlation of light upon coincidence – triggered Glauber's creation
of uniquely quantum coherence analysis. Classical optical coherence
becomes a classical limit for first-order quantum coherence; higher
degree of coherence leads to many phenomena in quantum optics.
Macroscopic quantum coherence
Macroscopic scale quantum coherence leads to novel phenomena, the so-called macroscopic quantum phenomena. For instance, the laser, superconductivity and superfluidity
are examples of highly coherent quantum systems whose effects are
evident at the macroscopic scale. The macroscopic quantum coherence
(off-diagonal long-range order, ODLRO)
for superfluidity, and laser light, is related to first-order (1-body)
coherence/ODLRO, while superconductivity is related to second-order
coherence/ODLRO. (For fermions, such as electrons, only even orders of
coherence/ODLRO are possible.) For bosons, a Bose–Einstein condensate is an example of a system exhibiting macroscopic quantum coherence through a multiple occupied single-particle state.
The classical electromagnetic field exhibits macroscopic quantum
coherence. The most obvious example is the carrier signal for radio and
TV. They satisfy Glauber's quantum description of coherence.
Quantum coherence as a resource
Recently M. B. Plenio
and co-workers constructed an operational formulation of quantum
coherence as a resource theory. They introduced coherence monotones
analogous to the entanglement monotones. Quantum coherence has been shown to be equivalent to quantum entanglement
in the sense that coherence can be faithfully described as
entanglement, and conversely that each entanglement measure corresponds
to a coherence measure.
Applications
Holography
Coherent superpositions of optical wave fields include holography. Holographic photographs have been used as art and as difficult to forge security labels.
Non-optical wave fields
Further applications concern the coherent superposition of non-optical wave fields. In quantum mechanics for example one considers a probability field, which is related to the wave function
(interpretation: density of the probability amplitude). Here the
applications concern, among others, the future technologies of quantum computing and the already available technology of quantum cryptography. Additionally the problems of the following subchapter are treated.
Modal analysis
Coherence
is used to check the quality of the transfer functions (FRFs) being
measured. Low coherence can be caused by poor signal to noise ratio,
and/or inadequate frequency resolution.