Quantum tunnelling

Last updated

In physics, quantum tunnelling, barrier penetration, or simply tunnelling is a quantum mechanical phenomenon in which an object such as an electron or atom passes through a potential energy barrier that, according to classical mechanics, should not be passable due to the object not having sufficient energy to pass or surmount the barrier.

Contents

Tunneling is a consequence of the wave nature of matter, where the quantum wave function describes the state of a particle or other physical system, and wave equations such as the Schrödinger equation describe their behavior. The probability of transmission of a wave packet through a barrier decreases exponentially with the barrier height, the barrier width, and the tunneling particle's mass, so tunneling is seen most prominently in low-mass particles such as electrons or protons tunneling through microscopically narrow barriers. Tunneling is readily detectable with barriers of thickness about 1–3 nm or smaller for electrons, and about 0.1 nm or smaller for heavier particles such as protons or hydrogen atoms. [1] Some sources describe the mere penetration of a wave function into the barrier, without transmission on the other side, as a tunneling effect, such as in tunneling into the walls of a finite potential well. [2] [3]

Tunneling plays an essential role in physical phenomena such as nuclear fusion [4] and alpha radioactive decay of atomic nuclei. Tunneling applications include the tunnel diode, [5] quantum computing, flash memory, and the scanning tunneling microscope. Tunneling limits the minimum size of devices used in microelectronics because electrons tunnel readily through insulating layers and transistors that are thinner than about 1 nm. [6]

The effect was predicted in the early 20th century. Its acceptance as a general physical phenomenon came mid-century. [7]

Introduction to the concept

Animation showing the tunnel effect and its application to an STM

Quantum tunnelling falls under the domain of quantum mechanics: the study of what happens at the quantum scale, which classical mechanics cannot explain. To understand the phenomenon, particles attempting to travel across a potential barrier can be compared to a ball trying to roll over a hill. Quantum mechanics and classical mechanics differ in their treatment of this scenario.

Classical mechanics predicts that particles that do not have enough energy to classically surmount a barrier cannot reach the other side. Thus, a ball without sufficient energy to surmount the hill would roll back down. In quantum mechanics, a particle can, with a small probability, tunnel to the other side, thus crossing the barrier. The reason for this difference comes from treating matter as having properties of waves and particles.

The tunnelling problem

A simulation of a wave packet incident on a potential barrier. In relative units, the barrier energy is 20, greater than the mean wave packet energy of 14. A portion of the wave packet passes through the barrier. E14-V20-B1.gif
A simulation of a wave packet incident on a potential barrier. In relative units, the barrier energy is 20, greater than the mean wave packet energy of 14. A portion of the wave packet passes through the barrier.

The wave function of a physical system of particles specifies everything that can be known about the system. [8] Therefore, problems in quantum mechanics analyze the system's wave function. Using mathematical formulations, such as the Schrödinger equation, the time evolution of a known wave function can be deduced. The square of the absolute value of this wave function is directly related to the probability distribution of the particle positions, which describes the probability that the particles would be measured at those positions.

As shown in the animation, a wave packet impinges on the barrier, most of it is reflected and some is transmitted through the barrier. The wave packet becomes more de-localized: it is now on both sides of the barrier and lower in maximum amplitude, but equal in integrated square-magnitude, meaning that the probability the particle is somewhere remains unity. The wider the barrier and the higher the barrier energy, the lower the probability of tunneling.

Some models of a tunneling barrier, such as the rectangular barriers shown, can be analysed and solved algebraically. [9] :96 Most problems do not have an algebraic solution, so numerical solutions are used. "Semiclassical methods" offer approximate solutions that are easier to compute, such as the WKB approximation.

History

The Schrödinger equation was published in 1926. The first person to apply the Schrödinger equation to a problem which involved tunneling between two classically allowed regions through a potential barrier was Friedrich Hund in a series of articles published in 1927. He studied the solutions of a double-well potential and discussed molecular spectra. [10] Leonid Mandelstam and Mikhail Leontovich discovered tunneling independently and published their results in 1928. [11]

In 1927, Lothar Nordheim, assisted by Ralph Fowler, published a paper which discussed thermionic emission and reflection of electrons from metals. He assumed a surface potential barrier which confines the electrons within the metal and showed that the electrons have a finite probability of tunneling through or reflecting from the surface barrier when their energies are close to the barrier energy. Classically, the electron would either transmit or reflect with 100% certainty, depending on its energy. In 1928 J. Robert Oppenheimer published two papers on field emission, i.e. the emission of electrons induced by strong electric fields. Nordheim and Fowler simplified Oppenheimer's derivation and found values for the emitted currents and work functions which agreed with experiments. [10]

A great success of the tunnelling theory was the mathematical explanation for alpha decay, which was developed in 1928 by George Gamow and independently by Ronald Gurney and Edward Condon. [12] [13] [14] [15] The latter researchers simultaneously solved the Schrödinger equation for a model nuclear potential and derived a relationship between the half-life of the particle and the energy of emission that depended directly on the mathematical probability of tunneling. All three researchers were familiar with the works on field emission, [10] and Gamow was aware of Mandelstam and Leontovich's findings. [16]

In the early days of quantum theory, the term tunnel effect was not used, and the effect was instead referred to as penetration of, or leaking through, a barrier. The German term wellenmechanische Tunneleffect was used in 1931 by Walter Schottky. The English term tunnel effect entered the language in 1932 when it was used by Yakov Frenkel in his textbook. [10]

In 1957 Leo Esaki demonstrated tunneling of electrons over a few nanometer wide barrier in a semiconductor structure and developed a diode based on tunnel effect. [17] In 1960, following Esaki's work, Ivar Giaever showed experimentally that tunnelling also took place in superconductors. The tunnelling spectrum gave direct evidence of the superconducting energy gap. In 1962, Brian Josephson predicted the tunneling of superconducting Cooper pairs. Esaki, Giaever and Josephson shared the 1973 Nobel Prize in Physics for their works on quantum tunneling in solids. [18] [7]

In 1981, Gerd Binnig and Heinrich Rohrer developed a new type of microscope, called scanning tunneling microscope, which is based on tunnelling and is used for imaging surfaces at the atomic level. Binnig and Rohrer were awarded the Nobel Prize in Physics in 1986 for their discovery. [19]

Applications

Tunnelling is the cause of some important macroscopic physical phenomena.

Solid-state physics

Electronics

Tunnelling is a source of current leakage in very-large-scale integration (VLSI) electronics and results in a substantial power drain and heating effects that plague such devices. It is considered the lower limit on how microelectronic device elements can be made. [20] Tunnelling is a fundamental technique used to program the floating gates of flash memory.

Cold emission

Cold emission of electrons is relevant to semiconductors and superconductor physics. It is similar to thermionic emission, where electrons randomly jump from the surface of a metal to follow a voltage bias because they statistically end up with more energy than the barrier, through random collisions with other particles. When the electric field is very large, the barrier becomes thin enough for electrons to tunnel out of the atomic state, leading to a current that varies approximately exponentially with the electric field. [21] These materials are important for flash memory, vacuum tubes, and some electron microscopes.

Tunnel junction

A simple barrier can be created by separating two conductors with a very thin insulator. These are tunnel junctions, the study of which requires understanding quantum tunnelling. [22] Josephson junctions take advantage of quantum tunnelling and superconductivity to create the Josephson effect. This has applications in precision measurements of voltages and magnetic fields, [21] as well as the multijunction solar cell.

Tunnel diode

A working mechanism of a resonant tunnelling diode device, based on the phenomenon of quantum tunnelling through the potential barriers Rtd seq v3.gif
A working mechanism of a resonant tunnelling diode device, based on the phenomenon of quantum tunnelling through the potential barriers

Diodes are electrical semiconductor devices that allow electric current flow in one direction more than the other. The device depends on a depletion layer between N-type and P-type semiconductors to serve its purpose. When these are heavily doped the depletion layer can be thin enough for tunnelling. When a small forward bias is applied, the current due to tunnelling is significant. This has a maximum at the point where the voltage bias is such that the energy level of the p and n conduction bands are the same. As the voltage bias is increased, the two conduction bands no longer line up and the diode acts typically. [23]

Because the tunnelling current drops off rapidly, tunnel diodes can be created that have a range of voltages for which current decreases as voltage increases. This peculiar property is used in some applications, such as high speed devices where the characteristic tunnelling probability changes as rapidly as the bias voltage. [23]

The resonant tunnelling diode makes use of quantum tunnelling in a very different manner to achieve a similar result. This diode has a resonant voltage for which a current favors a particular voltage, achieved by placing two thin layers with a high energy conductance band near each other. This creates a quantum potential well that has a discrete lowest energy level. When this energy level is higher than that of the electrons, no tunnelling occurs and the diode is in reverse bias. Once the two voltage energies align, the electrons flow like an open wire. As the voltage further increases, tunnelling becomes improbable and the diode acts like a normal diode again before a second energy level becomes noticeable. [24]

Tunnel field-effect transistors

A European research project demonstrated field effect transistors in which the gate (channel) is controlled via quantum tunnelling rather than by thermal injection, reducing gate voltage from ≈1 volt to 0.2 volts and reducing power consumption by up to 100×. If these transistors can be scaled up into VLSI chips, they would improve the performance per power of integrated circuits. [25] [26]

Conductivity of crystalline solids

While the Drude-Lorentz model of electrical conductivity makes excellent predictions about the nature of electrons conducting in metals, it can be furthered by using quantum tunnelling to explain the nature of the electron's collisions. [21] When a free electron wave packet encounters a long array of uniformly spaced barriers, the reflected part of the wave packet interferes uniformly with the transmitted one between all barriers so that 100% transmission becomes possible. The theory predicts that if positively charged nuclei form a perfectly rectangular array, electrons will tunnel through the metal as free electrons, leading to extremely high conductance, and that impurities in the metal will disrupt it. [21]

Scanning tunneling microscope

The scanning tunnelling microscope (STM), invented by Gerd Binnig and Heinrich Rohrer, may allow imaging of individual atoms on the surface of a material. [21] It operates by taking advantage of the relationship between quantum tunnelling with distance. When the tip of the STM's needle is brought close to a conduction surface that has a voltage bias, measuring the current of electrons that are tunnelling between the needle and the surface reveals the distance between the needle and the surface. By using piezoelectric rods that change in size when voltage is applied, the height of the tip can be adjusted to keep the tunnelling current constant. The time-varying voltages that are applied to these rods can be recorded and used to image the surface of the conductor. [21] STMs are accurate to 0.001 nm, or about 1% of atomic diameter. [24]

Nuclear physics

Nuclear fusion

Quantum tunnelling is an essential phenomenon for nuclear fusion. The temperature in stellar cores is generally insufficient to allow atomic nuclei to overcome the Coulomb barrier and achieve thermonuclear fusion. Quantum tunnelling increases the probability of penetrating this barrier. Though this probability is still low, the extremely large number of nuclei in the core of a star is sufficient to sustain a steady fusion reaction. [27]

Radioactive decay

Radioactive decay is the process of emission of particles and energy from the unstable nucleus of an atom to form a stable product. This is done via the tunnelling of a particle out of the nucleus (an electron tunneling into the nucleus is electron capture). This was the first application of quantum tunnelling. Radioactive decay is a relevant issue for astrobiology as this consequence of quantum tunnelling creates a constant energy source over a large time interval for environments outside the circumstellar habitable zone where insolation would not be possible (subsurface oceans) or effective. [27]

Quantum tunnelling may be one of the mechanisms of hypothetical proton decay. [28] [29]

Chemistry

Kinetic isotope effect

In chemical kinetics, the substitution of a light isotope of an element with a heavier one typically results in a slower reaction rate. This is generally attributed to differences in the zero-point vibrational energies for chemical bonds containing the lighter and heavier isotopes and is generally modeled using transition state theory. However, in certain cases, large isotopic effects are observed that cannot be accounted for by a semi-classical treatment, and quantum tunnelling is required. R. P. Bell developed a modified treatment of Arrhenius kinetics that is commonly used to model this phenomenon. [30]

Astrochemistry in interstellar clouds

By including quantum tunnelling, the astrochemical syntheses of various molecules in interstellar clouds can be explained, such as the synthesis of molecular hydrogen, water (ice) and the prebiotic important formaldehyde. [27] Tunnelling of molecular hydrogen has been observed in the lab. [31]

Quantum biology

Quantum tunnelling is among the central non-trivial quantum effects in quantum biology. [32] Here it is important both as electron tunnelling and proton tunnelling. Electron tunnelling is a key factor in many biochemical redox reactions (photosynthesis, cellular respiration) as well as enzymatic catalysis. Proton tunnelling is a key factor in spontaneous DNA mutation. [27]

Spontaneous mutation occurs when normal DNA replication takes place after a particularly significant proton has tunnelled. [33] A hydrogen bond joins DNA base pairs. A double well potential along a hydrogen bond separates a potential energy barrier. It is believed that the double well potential is asymmetric, with one well deeper than the other such that the proton normally rests in the deeper well. For a mutation to occur, the proton must have tunnelled into the shallower well. The proton's movement from its regular position is called a tautomeric transition. If DNA replication takes place in this state, the base pairing rule for DNA may be jeopardised, causing a mutation. [34] Per-Olov Lowdin was the first to develop this theory of spontaneous mutation within the double helix. Other instances of quantum tunnelling-induced mutations in biology are believed to be a cause of ageing and cancer. [35]

Mathematical discussion

Quantum tunnelling through a barrier. The energy of the tunnelled particle is the same but the probability amplitude is decreased. TunnelEffektKling1.png
Quantum tunnelling through a barrier. The energy of the tunnelled particle is the same but the probability amplitude is decreased.

The Schrödinger equation

The time-independent Schrödinger equation for one particle in one dimension can be written as

or

where

The solutions of the Schrödinger equation take different forms for different values of x, depending on whether M(x) is positive or negative. When M(x) is constant and negative, then the Schrödinger equation can be written in the form

The solutions of this equation represent travelling waves, with phase-constant +k or -k. Alternatively, if M(x) is constant and positive, then the Schrödinger equation can be written in the form

The solutions of this equation are rising and falling exponentials in the form of evanescent waves. When M(x) varies with position, the same difference in behaviour occurs, depending on whether M(x) is negative or positive. It follows that the sign of M(x) determines the nature of the medium, with negative M(x) corresponding to medium A and positive M(x) corresponding to medium B. It thus follows that evanescent wave coupling can occur if a region of positive M(x) is sandwiched between two regions of negative M(x), hence creating a potential barrier.

The mathematics of dealing with the situation where M(x) varies with x is difficult, except in special cases that usually do not correspond to physical reality. A full mathematical treatment appears in the 1965 monograph by Fröman and Fröman. Their ideas have not been incorporated into physics textbooks, but their corrections have little quantitative effect.

The WKB approximation

The wave function is expressed as the exponential of a function:

where

is then separated into real and imaginary parts:

where A(x) and B(x) are real-valued functions.

Substituting the second equation into the first and using the fact that the real part needs to be 0 results in:

Quantum tunneling in the phase space formulation of quantum mechanics. Wigner function for tunneling through the potential barrier in atomic units (a.u.). The solid lines represent the level set of the Hamiltonian .

To solve this equation using the semiclassical approximation, each function must be expanded as a power series in . From the equations, the power series must start with at least an order of to satisfy the real part of the equation; for a good classical limit starting with the highest power of Planck's constant possible is preferable, which leads to

and

with the following constraints on the lowest order terms,

and

At this point two extreme cases can be considered.

Case 1

If the amplitude varies slowly as compared to the phase and

which corresponds to classical motion. Resolving the next order of expansion yields

Case 2

If the phase varies slowly as compared to the amplitude, and

which corresponds to tunneling. Resolving the next order of the expansion yields

In both cases it is apparent from the denominator that both these approximate solutions are bad near the classical turning points . Away from the potential hill, the particle acts similar to a free and oscillating wave; beneath the potential hill, the particle undergoes exponential changes in amplitude. By considering the behaviour at these limits and classical turning points a global solution can be made.

To start, a classical turning point, is chosen and is expanded in a power series about :

Keeping only the first order term ensures linearity:

Using this approximation, the equation near becomes a differential equation:

This can be solved using Airy functions as solutions.

Taking these solutions for all classical turning points, a global solution can be formed that links the limiting solutions. Given the two coefficients on one side of a classical turning point, the two coefficients on the other side of a classical turning point can be determined by using this local solution to connect them.

Hence, the Airy function solutions will asymptote into sine, cosine and exponential functions in the proper limits. The relationships between and are

and

Quantum tunnelling through a barrier. At the origin (x = 0), there is a very high, but narrow potential barrier. A significant tunnelling effect can be seen. Quantum Tunnelling animation.gif
Quantum tunnelling through a barrier. At the origin (x = 0), there is a very high, but narrow potential barrier. A significant tunnelling effect can be seen.

With the coefficients found, the global solution can be found. Therefore, the transmission coefficient for a particle tunneling through a single potential barrier is

where are the two classical turning points for the potential barrier.

For a rectangular barrier, this expression simplifies to:

Faster than light

Some physicists have claimed that it is possible for spin-zero particles to travel faster than the speed of light when tunnelling. [7] This appears to violate the principle of causality, since a frame of reference then exists in which the particle arrives before it has left. In 1998, Francis E. Low reviewed briefly the phenomenon of zero-time tunnelling. [36] More recently, experimental tunnelling time data of phonons, photons, and electrons was published by Günter Nimtz. [37]

Other physicists, such as Herbert Winful, [38] disputed these claims. Winful argued that the wave packet of a tunnelling particle propagates locally, so a particle can't tunnel through the barrier non-locally. Winful also argued that the experiments that are purported to show non-local propagation have been misinterpreted. In particular, the group velocity of a wave packet does not measure its speed, but is related to the amount of time the wave packet is stored in the barrier. But the problem remains that the wave function still rises inside the barrier at all points at the same time. In other words, in any region that is inaccessible to measurement, non-local propagation is still mathematically certain.

A 2020 experiment, overseen by Aephraim M. Steinberg, showed that particles should be able to tunnel at apparent speeds faster than light. [39] [40]

Dynamical tunneling

Quantum tunneling oscillations of probability in an integrable double well of potential, seen in phase space Quantum tunneling in phase space.gif
Quantum tunneling oscillations of probability in an integrable double well of potential, seen in phase space

The concept of quantum tunneling can be extended to situations where there exists a quantum transport between regions that are classically not connected even if there is no associated potential barrier. This phenomenon is known as dynamical tunnelling. [41] [42]

Tunnelling in phase space

The concept of dynamical tunnelling is particularly suited to address the problem of quantum tunnelling in high dimensions (d>1). In the case of an integrable system, where bounded classical trajectories are confined onto tori in phase space, tunnelling can be understood as the quantum transport between semi-classical states built on two distinct but symmetric tori. [43]

Chaos-assisted tunnelling

Chaos-assisted tunnelling oscillations between two regular tori embedded in a chaotic sea, seen in phase space Chaos-assisted tunneling in phase space.gif
Chaos-assisted tunnelling oscillations between two regular tori embedded in a chaotic sea, seen in phase space

In real life, most systems are not integrable and display various degrees of chaos. Classical dynamics is then said to be mixed and the system phase space is typically composed of islands of regular orbits surrounded by a large sea of chaotic orbits. The existence of the chaotic sea, where transport is classically allowed, between the two symmetric tori then assists the quantum tunnelling between them. This phenomenon is referred as chaos-assisted tunnelling. [44] and is characterized by sharp resonances of the tunnelling rate when varying any system parameter.

Resonance-assisted tunnelling

When is small in front of the size of the regular islands, the fine structure of the classical phase space plays a key role in tunnelling. In particular the two symmetric tori are coupled "via a succession of classically forbidden transitions across nonlinear resonances" surrounding the two islands. [45]

Several phenomena have the same behavior as quantum tunnelling. Two examples are evanescent wave coupling [46] (the application of Maxwell's wave-equation to light) and the application of the non-dispersive wave-equation from acoustics applied to "waves on strings".[ citation needed ]

These effects are modeled similarly to the rectangular potential barrier. In these cases, one transmission medium through which the wave propagates that is the same or nearly the same throughout, and a second medium through which the wave travels differently. This can be described as a thin region of medium B between two regions of medium A. The analysis of a rectangular barrier by means of the Schrödinger equation can be adapted to these other effects provided that the wave equation has travelling wave solutions in medium A but real exponential solutions in medium B.

In optics, medium A is a vacuum while medium B is glass. In acoustics, medium A may be a liquid or gas and medium B a solid. For both cases, medium A is a region of space where the particle's total energy is greater than its potential energy and medium B is the potential barrier. These have an incoming wave and resultant waves in both directions. There can be more mediums and barriers, and the barriers need not be discrete. Approximations are useful in this case.

A classical wave-particle association was originally analyzed as analogous to quantum tunneling, [47] but subsequent analysis found a fluid dynamics cause related to the vertical momentum imparted to particles near the barrier. [48]

See also

Related Research Articles

<span class="mw-page-title-main">Particle in a box</span> Mathematical model in quantum mechanics

In quantum mechanics, the particle in a box model describes the movement of a free particle in a small space surrounded by impenetrable barriers. The model is mainly used as a hypothetical example to illustrate the differences between classical and quantum systems. In classical systems, for example, a particle trapped inside a large box can move at any speed within the box and it is no more likely to be found at one position than another. However, when the well becomes very narrow, quantum effects become important. The particle may only occupy certain positive energy levels. Likewise, it can never have zero energy, meaning that the particle can never "sit still". Additionally, it is more likely to be found at certain positions than at others, depending on its energy level. The particle may never be detected at certain positions, known as spatial nodes.

<span class="mw-page-title-main">Quantum mechanics</span> Description of physical properties at the atomic and subatomic scale

Quantum mechanics is a fundamental theory in physics that describes the behavior of nature at and below the scale of atoms. It is the foundation of all quantum physics, which includes quantum chemistry, quantum field theory, quantum technology, and quantum information science.

<span class="mw-page-title-main">Scanning tunneling microscope</span> Instrument able to image surfaces at the atomic level by exploiting quantum tunneling effects

A scanning tunneling microscope (STM) is a type of microscope used for imaging surfaces at the atomic level. Its development in 1981 earned its inventors, Gerd Binnig and Heinrich Rohrer, then at IBM Zürich, the Nobel Prize in Physics in 1986. STM senses the surface by using an extremely sharp conducting tip that can distinguish features smaller than 0.1 nm with a 0.01 nm (10 pm) depth resolution. This means that individual atoms can routinely be imaged and manipulated. Most scanning tunneling microscopes are built for use in ultra-high vacuum at temperatures approaching absolute zero, but variants exist for studies in air, water and other environments, and for temperatures over 1000 °C.

<span class="mw-page-title-main">Schrödinger equation</span> Description of a quantum-mechanical system

The Schrödinger equation is a linear partial differential equation that governs the wave function of a quantum-mechanical system. Its discovery was a significant landmark in the development of quantum mechanics. It is named after Erwin Schrödinger, who postulated the equation in 1925 and published it in 1926, forming the basis for the work that resulted in his Nobel Prize in Physics in 1933.

The Klein–Gordon equation is a relativistic wave equation, related to the Schrödinger equation. It is second-order in space and time and manifestly Lorentz-covariant. It is a differential equation version of the relativistic energy–momentum relation .

In quantum mechanics, the particle in a one-dimensional lattice is a problem that occurs in the model of a periodic crystal lattice. The potential is caused by ions in the periodic structure of the crystal creating an electromagnetic field so electrons are subject to a regular potential inside the lattice. It is a generalization of the free electron model, which assumes zero potential inside the lattice.

<span class="mw-page-title-main">Relativistic wave equations</span> Wave equations respecting special and general relativity

In physics, specifically relativistic quantum mechanics (RQM) and its applications to particle physics, relativistic wave equations predict the behavior of particles at high energies and velocities comparable to the speed of light. In the context of quantum field theory (QFT), the equations determine the dynamics of quantum fields. The solutions to the equations, universally denoted as ψ or Ψ, are referred to as "wave functions" in the context of RQM, and "fields" in the context of QFT. The equations themselves are called "wave equations" or "field equations", because they have the mathematical form of a wave equation or are generated from a Lagrangian density and the field-theoretic Euler–Lagrange equations.

<span class="mw-page-title-main">Pilot wave theory</span> One interpretation of quantum mechanics

In theoretical physics, the pilot wave theory, also known as Bohmian mechanics, was the first known example of a hidden-variable theory, presented by Louis de Broglie in 1927. Its more modern version, the de Broglie–Bohm theory, interprets quantum mechanics as a deterministic theory, and avoids issues such as wave–particle duality, instantaneous wave function collapse, and the paradox of Schrödinger's cat by being inherently nonlocal.

In physics, a free particle is a particle that, in some sense, is not bound by an external force, or equivalently not in a region where its potential energy varies. In classical physics, this means the particle is present in a "field-free" space. In quantum mechanics, it means the particle is in a region of uniform potential, usually set to zero in the region of interest since the potential can be arbitrarily set to zero at any point in space.

The Schrödinger–Newton equation, sometimes referred to as the Newton–Schrödinger or Schrödinger–Poisson equation, is a nonlinear modification of the Schrödinger equation with a Newtonian gravitational potential, where the gravitational potential emerges from the treatment of the wave function as a mass density, including a term that represents interaction of a particle with its own gravitational field. The inclusion of a self-interaction term represents a fundamental alteration of quantum mechanics. It can be written either as a single integro-differential equation or as a coupled system of a Schrödinger and a Poisson equation. In the latter case it is also referred to in the plural form.

In quantum mechanics and scattering theory, the one-dimensional step potential is an idealized system used to model incident, reflected and transmitted matter waves. The problem consists of solving the time-independent Schrödinger equation for a particle with a step-like potential in one dimension. Typically, the potential is modeled as a Heaviside step function.

<span class="mw-page-title-main">Rectangular potential barrier</span> Area, where a potential exhibits a local maximum

In quantum mechanics, the rectangularpotential barrier is a standard one-dimensional problem that demonstrates the phenomena of wave-mechanical tunneling and wave-mechanical reflection. The problem consists of solving the one-dimensional time-independent Schrödinger equation for a particle encountering a rectangular potential energy barrier. It is usually assumed, as here, that a free particle impinges on the barrier from the left.

In quantum mechanics the delta potential is a potential well mathematically described by the Dirac delta function - a generalized function. Qualitatively, it corresponds to a potential which is zero everywhere, except at a single point, where it takes an infinite value. This can be used to simulate situations where a particle is free to move in two regions of space with a barrier between the two regions. For example, an electron can move almost freely in a conducting material, but if two conducting surfaces are put close together, the interface between them acts as a barrier for the electron that can be approximated by a delta potential.

The theoretical and experimental justification for the Schrödinger equation motivates the discovery of the Schrödinger equation, the equation that describes the dynamics of nonrelativistic particles. The motivation uses photons, which are relativistic particles with dynamics described by Maxwell's equations, as an analogue for all types of particles.

The quantum potential or quantum potentiality is a central concept of the de Broglie–Bohm formulation of quantum mechanics, introduced by David Bohm in 1952.

The Gamow factor, Sommerfeld factor or Gamow–Sommerfeld factor, named after its discoverer George Gamow or after Arnold Sommerfeld, is a probability factor for two nuclear particles' chance of overcoming the Coulomb barrier in order to undergo nuclear reactions, for example in nuclear fusion. By classical physics, there is almost no possibility for protons to fuse by crossing each other's Coulomb barrier at temperatures commonly observed to cause fusion, such as those found in the sun. When George Gamow instead applied quantum mechanics to the problem, he found that there was a significant chance for the fusion due to tunneling.

In physics, relativistic quantum mechanics (RQM) is any Poincaré covariant formulation of quantum mechanics (QM). This theory is applicable to massive particles propagating at all velocities up to those comparable to the speed of light c, and can accommodate massless particles. The theory has application in high energy physics, particle physics and accelerator physics, as well as atomic physics, chemistry and condensed matter physics. Non-relativistic quantum mechanics refers to the mathematical formulation of quantum mechanics applied in the context of Galilean relativity, more specifically quantizing the equations of classical mechanics by replacing dynamical variables by operators. Relativistic quantum mechanics (RQM) is quantum mechanics applied with special relativity. Although the earlier formulations, like the Schrödinger picture and Heisenberg picture were originally formulated in a non-relativistic background, a few of them also work with special relativity.

In quantum mechanics, energy is defined in terms of the energy operator, acting on the wave function of the system as a consequence of time translation symmetry.

In quantum mechanics, resonance cross section occurs in the context of quantum scattering theory, which deals with studying the scattering of quantum particles from potentials. The scattering problem deals with the calculation of flux distribution of scattered particles/waves as a function of the potential, and of the state of the incident particle. For a free quantum particle incident on the potential, the plane wave solution to the time-independent Schrödinger wave equation is:

References

  1. Lerner; Trigg (1991). Encyclopedia of Physics (2nd ed.). New York: VCH. p.  1308. ISBN   978-0-89573-752-6.
  2. Davies, P C W (6 May 2004). "Quantum mechanics and the equivalence principle". Classical and Quantum Gravity. 21 (11): 2761–2772. arXiv: quant-ph/0403027 . doi:10.1088/0264-9381/21/11/017. ISSN   0264-9381. But quantum particles are able to tunnel into the classically forbidden region ...
  3. Fowler, Michael. "Particle in a Finite Box and Tunneling". LibreTexts Chemistry. Retrieved 4 September 2023. Tunneling into the barrier (wall) is possible.
  4. Serway; Vuille (2008). College Physics. Vol. 2 (Eighth ed.). Belmont: Brooks/Cole. ISBN   978-0-495-55475-2.
  5. Taylor, J. (2004). Modern Physics for Scientists and Engineers. Prentice Hall. p. 234. ISBN   978-0-13-805715-2.
  6. "Quantum Effects At 7/5nm And Beyond". Semiconductor Engineering. Retrieved 15 July 2018.
  7. 1 2 3 Razavy, Mohsen (2003). Quantum Theory of Tunneling . World Scientific. pp.  4, 462. ISBN   978-9812564887.
  8. Bjorken and Drell, Relativistic Quantum Mechanics, Mcgraw-Hill College, 1965. p. 2
  9. Messiah, Albert (1966). Quantum Mechanics. North Holland, John Wiley & Sons. ISBN   0486409244.
  10. 1 2 3 4 Merzbacher, Eugen (August 2002). "The Early History of Quantum Tunneling". Physics Today. 55 (8): 44–49. Bibcode:2002PhT....55h..44M. doi:10.1063/1.1510281 . Retrieved 17 August 2022. Friedrich Hund ... was the first to make use of quantum mechanical barrier penetration ...
  11. Mandelstam, L.; Leontowitsch, M. (1928). "Zur Theorie der Schrödingerschen Gleichung". Zeitschrift für Physik. 47 (1–2): 131–136. Bibcode:1928ZPhy...47..131M. doi:10.1007/BF01391061. S2CID   125101370.
  12. Gurney, R. W.; Condon, E. U. (1928). "Quantum Mechanics and Radioactive Disintegration". Nature. 122 (3073): 439. Bibcode:1928Natur.122..439G. doi: 10.1038/122439a0 . S2CID   4090561.
  13. Gurney, R. W.; Condon, E. U. (1929). "Quantum Mechanics and Radioactive Disintegration". Physical Review. 33 (2): 127–140. Bibcode:1929PhRv...33..127G. doi:10.1103/PhysRev.33.127.
  14. Bethe, Hans (27 October 1966). "Hans Bethe – Session I". Niels Bohr Library & Archives, American Institute of Physics, College Park, Maryland, USA (Interview). Interviewed by Charles Weiner; Jagdish Mehra. Cornell University. Retrieved 1 May 2016.
  15. Friedlander, Gerhart; Kennedy, Joseph E.; Miller, Julian Malcolm (1964). Nuclear and Radiochemistry (2nd ed.). New York: John Wiley & Sons. pp.  225–7. ISBN   978-0-471-86255-0.
  16. Feinberg, E. L. (2002). "The forefather (about Leonid Isaakovich Mandelstam)". Physics-Uspekhi. 45 (1): 81–100. Bibcode:2002PhyU...45...81F. doi:10.1070/PU2002v045n01ABEH001126. S2CID   250780246.
  17. Esaki, Leo (22 March 1974). "Long Journey into Tunneling". Science. 183 (4130): 1149–1155. Bibcode:1974Sci...183.1149E. doi:10.1126/science.183.4130.1149. ISSN   0036-8075. PMID   17789212. S2CID   44642243.
  18. Dardo, M. (Mauro) (2004). Nobel laureates and twentieth-century physics. Internet Archive. Cambridge, UK ; New York : Cambridge University Press. ISBN   978-0-521-83247-2.
  19. Binnig G, Rohrer H (1 July 1987). "Scanning tunneling microscopy—from birth to adolescence". Reviews of Modern Physics. 59 (3): 615–625. Bibcode:1987RvMP...59..615B. doi: 10.1103/RevModPhys.59.615 .
  20. "Applications of tunneling". psi.phys.wits.ac.za. Retrieved 30 April 2023.
  21. 1 2 3 4 5 6 Taylor, J. (2004). Modern Physics for Scientists and Engineers. Prentice Hall. p. 479. ISBN   978-0-13-805715-2.
  22. Lerner; Trigg (1991). Encyclopedia of Physics (2nd ed.). New York: VCH. pp.  1308–1309. ISBN   978-0-89573-752-6.
  23. 1 2 Krane, Kenneth (1983). Modern Physics. New York: John Wiley and Sons. p.  423. ISBN   978-0-471-07963-7.
  24. 1 2 Knight, R. D. (2004). Physics for Scientists and Engineers: With Modern Physics. Pearson Education. p. 1311. ISBN   978-0-321-22369-2.
  25. Ionescu, Adrian M.; Riel, Heike (2011). "Tunnel field-effect transistors as energy-efficient electronic switches". Nature . 479 (7373): 329–337. Bibcode:2011Natur.479..329I. doi:10.1038/nature10679. PMID   22094693. S2CID   4322368.
  26. Vyas, P. B.; Naquin, C.; Edwards, H.; Lee, M.; Vandenberghe, W. G.; Fischetti, M. V. (23 January 2017). "Theoretical simulation of negative differential transconductance in lateral quantum well nMOS devices". Journal of Applied Physics. 121 (4): 044501. Bibcode:2017JAP...121d4501V. doi:10.1063/1.4974469. ISSN   0021-8979.
  27. 1 2 3 4 Trixler, F. (2013). "Quantum tunnelling to the origin and evolution of life". Current Organic Chemistry. 17 (16): 1758–1770. doi:10.2174/13852728113179990083. PMC   3768233 . PMID   24039543.
  28. Talou, P.; Carjan, N.; Strottman, D. (1998). "Time-dependent properties of proton decay from crossing single-particle metastable states in deformed nuclei". Physical Review C. 58 (6): 3280–3285. arXiv: nucl-th/9809006 . Bibcode:1998PhRvC..58.3280T. doi:10.1103/PhysRevC.58.3280. S2CID   119075457.
  29. "adsabs.harvard.edu".
  30. Bell, Ronald Percy (1980). The tunnel effect in chemistry. London: Chapman and Hall. ISBN   0412213400. OCLC   6854792.
  31. Wild, Robert; Nötzold, Markus; Simpson, Malcolm; Tran, Thuy Dung; Wester, Roland (1 March 2023). "Tunnelling measured in a very slow ion–molecule reaction". Nature. 615 (7952): 425–429. arXiv: 2303.14948 . Bibcode:2023Natur.615..425W. doi:10.1038/s41586-023-05727-z. ISSN   1476-4687. PMID   36859549. S2CID   257282176.
  32. Trixler, F. (2013). "Quantum Tunnelling to the Origin and Evolution of Life". Current Organic Chemistry. 17 (16): 1758–1770. doi:10.2174/13852728113179990083. PMC   3768233 . PMID   24039543.
  33. Matta, Cherif F. (2014). Quantum Biochemistry: Electronic Structure and Biological Activity. Weinheim: Wiley-VCH. ISBN   978-3-527-62922-0.
  34. Majumdar, Rabi (2011). Quantum Mechanics: In Physics and Chemistry with Applications to Bioloty. Newi: PHI Learning. ISBN   9788120343047.
  35. Cooper, W. G. (June 1993). "Roles of Evolution, Quantum Mechanics and Point Mutations in Origins of Cancer". Cancer Biochemistry Biophysics. 13 (3): 147–170. PMID   8111728.
  36. Low, F. E. (1998). "Comments on apparent superluminal propagation". Ann. Phys. 7 (7–8): 660–661. Bibcode:1998AnP...510..660L. doi:10.1002/(SICI)1521-3889(199812)7:7/8<660::AID-ANDP660>3.0.CO;2-0. S2CID   122717505.
  37. Nimtz, G. (2011). "Tunneling Confronts Special Relativity". Found. Phys. 41 (7): 1193–1199. arXiv: 1003.3944 . Bibcode:2011FoPh...41.1193N. doi:10.1007/s10701-011-9539-2. S2CID   119249900.
  38. Winful, H. G. (2006). "Tunneling time, the Hartman effect, and superluminality: A proposed resolution of an old paradox". Phys. Rep. 436 (1–2): 1–69. Bibcode:2006PhR...436....1W. doi:10.1016/j.physrep.2006.09.002.
  39. "Quantum-tunnelling time is measured using ultracold atoms – Physics World". 22 July 2020.
  40. "Quanta Magazine". 20 October 2020.
  41. Davis, Michael J.; Heller, Eric J. (1 July 1981). "Quantum dynamical tunneling in bound states". The Journal of Chemical Physics. 75 (1): 246–254. Bibcode:1981JChPh..75..246D. doi:10.1063/1.441832. ISSN   0021-9606.
  42. Keshavamurthy, Srihari; Schlagheck, Peter (9 March 2011). Dynamical Tunneling: Theory and Experiment. CRC Press. ISBN   978-1-4398-1666-0.
  43. Wilkinson, Michael (1 September 1986). "Tunnelling between tori in phase space". Physica D: Nonlinear Phenomena. 21 (2): 341–354. Bibcode:1986PhyD...21..341W. doi:10.1016/0167-2789(86)90009-6. ISSN   0167-2789.
  44. Tomsovic, Steven; Ullmo, Denis (1 July 1994). "Chaos-assisted tunneling". Physical Review E. 50 (1): 145–162. Bibcode:1994PhRvE..50..145T. doi:10.1103/PhysRevE.50.145. PMID   9961952.
  45. Brodier, Olivier; Schlagheck, Peter; Ullmo, Denis (25 August 2002). "Resonance-Assisted Tunneling". Annals of Physics. 300 (1): 88–136. arXiv: nlin/0205054 . Bibcode:2002AnPhy.300...88B. doi:10.1006/aphy.2002.6281. ISSN   0003-4916. S2CID   51895893.
  46. Martin, Th.; Landauer, R. (1 February 1992). "Time delay of evanescent electromagnetic waves and the analogy to particle tunneling". Physical Review A. 45 (4): 2611–2617. doi:10.1103/PhysRevA.45.2611. ISSN   1050-2947.
  47. Eddi, A.; Fort, E.; Moisy, F.; Couder, Y. (16 June 2009). "Unpredictable Tunneling of a Classical Wave-Particle Association" (PDF). Physical Review Letters. 102 (24): 240401. Bibcode:2009PhRvL.102x0401E. doi:10.1103/PhysRevLett.102.240401. PMID   19658983. Archived from the original (PDF) on 10 March 2016. Retrieved 1 May 2016.
  48. Bush, John W M; Oza, Anand U (1 January 2021). "Hydrodynamic quantum analogs". Reports on Progress in Physics. 84 (1): 017001. doi:10.1088/1361-6633/abc22c. ISSN   0034-4885.

Further reading