Yukawa potential

Last updated

In particle, atomic and condensed matter physics, a Yukawa potential (also called a screened Coulomb potential) is a potential named after the Japanese physicist Hideki Yukawa. The potential is of the form:

Contents

where g is a magnitude scaling constant, i.e. is the amplitude of potential, m is the mass of the particle, r is the radial distance to the particle, and α is another scaling constant, so that is the approximate range. The potential is monotonically increasing in r and it is negative, implying the force is attractive. In the SI system, the unit of the Yukawa potential is (1/meters).

The Coulomb potential of electromagnetism is an example of a Yukawa potential with the factor equal to 1, everywhere. This can be interpreted as saying that the photon mass m is equal to 0. The photon is the force-carrier between interacting, charged particles.

In interactions between a meson field and a fermion field, the constant g is equal to the gauge coupling constant between those fields. In the case of the nuclear force, the fermions would be a proton and another proton or a neutron.

History

Prior to Hideki Yukawa's 1935 paper, [1] physicists struggled to explain the results of James Chadwick's atomic model, which consisted of positively charged protons and neutrons packed inside of a small nucleus, with a radius on the order of 10−14 meters. Physicists knew that electromagnetic forces at these lengths would cause these protons to repel each other and for the nucleus to fall apart. [2] Thus came the motivation for further explaining the interactions between elementary particles. In 1932, Werner Heisenberg proposed a "Platzwechsel" (migration) interaction between the neutrons and protons inside the nucleus, in which neutrons were composite particles of protons and electrons. These composite neutrons would emit electrons, creating an attractive force with the protons, and then turn into protons themselves. When, in 1933 at the Solvay Conference, Heisenberg proposed his interaction, physicists suspected it to be of either two forms:

on account of its short-range. [3] However, there were many issues with his theory. Namely, it is impossible for an electron of spin 1/2 and a proton of spin 1/2 to add up to the neutron spin of 1/2. The way Heisenberg treated this issue would go on to form the ideas of isospin.

Heisenberg's idea of an exchange interaction (rather than a Coulombic force) between particles inside the nucleus led Fermi to formulate his ideas on beta-decay in 1934. [3] Fermi's neutron-proton interaction was not based on the "migration" of neutron and protons between each other. Instead, Fermi proposed the emission and absorption of two light particles: the neutrino and electron, rather than just the electron (as in Heisenberg's theory). While Fermi's interaction solved the issue of the conservation of linear and angular momentum, Soviet physicists Igor Tamm and Dmitri Ivanenko demonstrated that the force associated with the neutrino and electron emission was not strong enough to bind the protons and neutrons in the nucleus. [4]

In his February 1935 paper, Hideki Yukawa combines both the idea of Heisenberg's short-range force interaction and Fermi's idea of an exchange particle in order to fix the issue of the neutron-proton interaction. He deduced a potential which includes an exponential decay term () and an electromagnetic term (). In analogy to quantum field theory, Yukawa knew that the potential and its corresponding field must be a result of an exchange particle. In the case of QED, this exchange particle was a photon of 0 mass. In Yukawa's case, the exchange particle had some mass, which was related to the range of interaction (given by ). Since the range of the nuclear force was known, Yukawa used his equation to predict the mass of the mediating particle as about 200 times the mass of the electron. Physicists called this particle the "meson," as its mass was in the middle of the proton and electron. Yukawa's meson was found in 1947, and came to be known as the pion. [4]

Relation to Coulomb potential

Figure 1: A comparison of Yukawa potentials where g = 1 and with various values for m. Yukawa m compare.svg
Figure 1: A comparison of Yukawa potentials where g = 1 and with various values for m.
Figure 2: A "long-range" comparison of Yukawa and Coulomb potentials' strengths where g = 1. Yukawa coulomb compare.svg
Figure 2: A "long-range" comparison of Yukawa and Coulomb potentials' strengths where g = 1.

If the particle has no mass (i.e., m = 0), then the Yukawa potential reduces to a Coulomb potential, and the range is said to be infinite. In fact, we have:

Consequently, the equation

simplifies to the form of the Coulomb potential

where we set the scaling constant to be: [5]

A comparison of the long range potential strength for Yukawa and Coulomb is shown in Figure 2. It can be seen that the Coulomb potential has effect over a greater distance whereas the Yukawa potential approaches zero rather quickly. However, any Yukawa potential or Coulomb potential is non-zero for any large r.

Fourier transform

The easiest way to understand that the Yukawa potential is associated with a massive field is by examining its Fourier transform. One has

where the integral is performed over all possible values of the 3-vector momenta k. In this form, and setting the scaling factor to one, , the fraction is seen to be the propagator or Green's function of the Klein–Gordon equation.

Feynman amplitude

Single particle exchange. Momentum exchange.svg
Single particle exchange.

The Yukawa potential can be derived as the lowest order amplitude of the interaction of a pair of fermions. The Yukawa interaction couples the fermion field to the meson field with the coupling term

The scattering amplitude for two fermions, one with initial momentum and the other with momentum , exchanging a meson with momentum k, is given by the Feynman diagram on the right.

The Feynman rules for each vertex associate a factor of g with the amplitude; since this diagram has two vertices, the total amplitude will have a factor of . The line in the middle, connecting the two fermion lines, represents the exchange of a meson. The Feynman rule for a particle exchange is to use the propagator; the propagator for a massive meson is . Thus, we see that the Feynman amplitude for this graph is nothing more than

From the previous section, this is seen to be the Fourier transform of the Yukawa potential.

Eigenvalues of Schrödinger equation

The radial Schrödinger equation with Yukawa potential can be solved perturbatively. [6] [7] [8] :ch. 16 Using the radial Schrödinger equation in the form

and the Yukawa potential in the power-expanded form

and setting , one obtains for the angular momentum the expression

for , where

Setting all coefficients except equal to zero, one obtains the well-known expression for the Schrödinger eigenvalue for the Coulomb potential, and the radial quantum number is a positive integer or zero as a consequence of the boundary conditions which the wave functions of the Coulomb potential have to satisfy. In the case of the Yukawa potential the imposition of boundary conditions is more complicated. Thus in the Yukawa case is only an approximation and the parameter that replaces the integer n is really an asymptotic expansion like that above with first approximation the integer value of the corresponding Coulomb case. The above expansion for the orbital angular momentum or Regge trajectory can be reversed to obtain the energy eigenvalues or equivalently . One obtains: [9]

The above asymptotic expansion of the angular momentum in descending powers of can also be derived with the WKB method. In that case, however, as in the case of the Coulomb potential the expression in the centrifugal term of the Schrödinger equation has to be replaced by , as was argued originally by Langer, [10] the reason being that the singularity is too strong for an unchanged application of the WKB method. That this reasoning is correct follows from the WKB derivation of the correct result in the Coulomb case (with the Langer correction), [8] :404 and even of the above expansion in the Yukawa case with higher order WKB approximations. [11]

Cross section

We can calculate the differential cross section between a proton or neutron and the pion by making use of the Yukawa potential. We use the Born approximation, which tells us that, in a spherically symmetrical potential, we can approximate the outgoing scattered wave function as the sum of incoming plane wave function and a small perturbation:

where is the particle's incoming momentum. The function is given by:

where is the particle's outgoing scattered momentum and is the incoming particles' mass (not to be confused with the pion's mass). We calculate by plugging in :

Evaluating the integral gives

Energy conservation implies

so that

Plugging in, we get:

We thus get a differential cross section of: [5]

Integrating, the total cross section is:

See also

Related Research Articles

<span class="mw-page-title-main">Arithmetic–geometric mean</span> Mathematical function of two positive real arguments

In mathematics, the arithmetic–geometric mean of two positive real numbers x and y is the mutual limit of a sequence of arithmetic means and a sequence of geometric means:

In physics, the cross section is a measure of the probability that a specific process will take place when some kind of radiant excitation intersects a localized phenomenon. For example, the Rutherford cross-section is a measure of probability that an alpha particle will be deflected by a given angle during an interaction with an atomic nucleus. Cross section is typically denoted σ (sigma) and is expressed in units of area, more specifically in barns. In a way, it can be thought of as the size of the object that the excitation must hit in order for the process to occur, but more exactly, it is a parameter of a stochastic process.

In integral calculus, an elliptic integral is one of a number of related functions defined as the value of certain integrals, which were first studied by Giulio Fagnano and Leonhard Euler. Their name originates from their originally arising in connection with the problem of finding the arc length of an ellipse.

<span class="mw-page-title-main">Electroweak interaction</span> Unified description of electromagnetism and the weak interaction

In particle physics, the electroweak interaction or electroweak force is the unified description of two of the four known fundamental interactions of nature: electromagnetism and the weak interaction. Although these two forces appear very different at everyday low energies, the theory models them as two different aspects of the same force. Above the unification energy, on the order of 246 GeV, they would merge into a single force. Thus, if the temperature is high enough – approximately 1015 K – then the electromagnetic force and weak force merge into a combined electroweak force. During the quark epoch (shortly after the Big Bang), the electroweak force split into the electromagnetic and weak force. It is thought that the required temperature of 1015 K has not been seen widely throughout the universe since before the quark epoch, and currently the highest man-made temperature in thermal equilibrium is around 5.5x1012 K (from the Large Hadron Collider).

<span class="mw-page-title-main">Hydrogen atom</span> Atom of the element hydrogen

A hydrogen atom is an atom of the chemical element hydrogen. The electrically neutral atom contains a single positively charged proton and a single negatively charged electron bound to the nucleus by the Coulomb force. Atomic hydrogen constitutes about 75% of the baryonic mass of the universe.

<span class="mw-page-title-main">Pauli matrices</span> Matrices important in quantum mechanics and the study of spin

In mathematical physics and mathematics, the Pauli matrices are a set of three 2 × 2 complex matrices which are Hermitian, involutory and unitary. Usually indicated by the Greek letter sigma, they are occasionally denoted by tau when used in connection with isospin symmetries.

<span class="mw-page-title-main">Rutherford scattering</span> Elastic scattering of charged particles by the Coulomb force

In particle physics, Rutherford scattering is the elastic scattering of charged particles by the Coulomb interaction. It is a physical phenomenon explained by Ernest Rutherford in 1911 that led to the development of the planetary Rutherford model of the atom and eventually the Bohr model. Rutherford scattering was first referred to as Coulomb scattering because it relies only upon the static electric (Coulomb) potential, and the minimum distance between particles is set entirely by this potential. The classical Rutherford scattering process of alpha particles against gold nuclei is an example of "elastic scattering" because neither the alpha particles nor the gold nuclei are internally excited. The Rutherford formula further neglects the recoil kinetic energy of the massive target nucleus.

<span class="mw-page-title-main">Legendre polynomials</span> System of complete and orthogonal polynomials

In mathematics, Legendre polynomials, named after Adrien-Marie Legendre (1782), are a system of complete and orthogonal polynomials with a vast number of mathematical properties and numerous applications. They can be defined in many ways, and the various definitions highlight different aspects as well as suggest generalizations and connections to different mathematical structures and physical and numerical applications.

In statistics, maximum likelihood estimation (MLE) is a method of estimating the parameters of an assumed probability distribution, given some observed data. This is achieved by maximizing a likelihood function so that, under the assumed statistical model, the observed data is most probable. The point in the parameter space that maximizes the likelihood function is called the maximum likelihood estimate. The logic of maximum likelihood is both intuitive and flexible, and as such the method has become a dominant means of statistical inference.

In mathematical analysis, Hölder's inequality, named after Otto Hölder, is a fundamental inequality between integrals and an indispensable tool for the study of Lp spaces.

In mathematics, especially vector calculus and differential topology, a closed form is a differential form α whose exterior derivative is zero, and an exact form is a differential form, α, that is the exterior derivative of another differential form β. Thus, an exact form is in the image of d, and a closed form is in the kernel of d.

In quantum mechanics, a particle in a spherically symmetric potential, is a quantum system with a potential that depends only on the distance between the particle and a defined center point. One example of a spherically symmetric potential is the electron within a hydrogen atom. The electron's potential only depends on its distance from the proton in the atom's nucleus. This potential can be derived from Coulomb's law.

<span class="mw-page-title-main">Lamb shift</span> Difference in energy of hydrogenic atom electron states not predicted by the Dirac equation

In physics, the Lamb shift, named after Willis Lamb, is a difference in energy between two energy levels 2S1/2 and 2P1/2 of the hydrogen atom which was not predicted by the Dirac equation, according to which these states should have the same energy.

In particle physics, Yukawa's interaction or Yukawa coupling, named after Hideki Yukawa, is an interaction between particles according to the Yukawa potential. Specifically, it is a scalar field ϕ and a Dirac field ψ of the type

A multipole expansion is a mathematical series representing a function that depends on angles—usually the two angles used in the spherical coordinate system for three-dimensional Euclidean space, . Similarly to Taylor series, multipole expansions are useful because oftentimes only the first few terms are needed to provide a good approximation of the original function. The function being expanded may be real- or complex-valued and is defined either on , or less often on for some other .

A hydrogen-like atom (or hydrogenic atom) is any atom or ion with a single valence electron. These atoms are isoelectronic with hydrogen. Examples of hydrogen-like atoms include, but are not limited to, hydrogen itself, all alkali metals such as Rb and Cs, singly ionized alkaline earth metals such as Ca+ and Sr+ and other ions such as He+, Li2+, and Be3+ and isotopes of any of the above. A hydrogen-like atom includes a positively charged core consisting of the atomic nucleus and any core electrons as well as a single valence electron. Because helium is common in the universe, the spectroscopy of singly ionized helium is important in EUV astronomy, for example, of DO white dwarf stars.

<span class="mw-page-title-main">Inverse curve</span> Curve created by a geometric operation

In inversive geometry, an inverse curve of a given curve C is the result of applying an inverse operation to C. Specifically, with respect to a fixed circle with center O and radius k the inverse of a point Q is the point P for which P lies on the ray OQ and OP·OQ = k2. The inverse of the curve C is then the locus of P as Q runs over C. The point O in this construction is called the center of inversion, the circle the circle of inversion, and k the radius of inversion.

<span class="mw-page-title-main">Coulomb wave function</span>

In mathematics, a Coulomb wave function is a solution of the Coulomb wave equation, named after Charles-Augustin de Coulomb. They are used to describe the behavior of charged particles in a Coulomb potential and can be written in terms of confluent hypergeometric functions or Whittaker functions of imaginary argument.

Static force fields are fields, such as a simple electric, magnetic or gravitational fields, that exist without excitations. The most common approximation method that physicists use for scattering calculations can be interpreted as static forces arising from the interactions between two bodies mediated by virtual particles, particles that exist for only a short time determined by the uncertainty principle. The virtual particles, also known as force carriers, are bosons, with different bosons associated with each force.

Partial-wave analysis, in the context of quantum mechanics, refers to a technique for solving scattering problems by decomposing each wave into its constituent angular-momentum components and solving using boundary conditions.

References

  1. Yukawa, H. (1935). "On the interaction of elementary particles". Proc. Phys.-Math. Soc. Jpn. 17: 48.
  2. Lincoln, Don (2004). Understanding the Universe: From quarks to the cosmos . Singapore: World Scientific. pp.  75–78. ISBN   978-9812387035.
  3. 1 2 Miller, Arthur I. (1985). "Werner Heisenberg and the beginning of nuclear physics". Physics Today. 38 (11): 60–68. Bibcode:1985PhT....38k..60M. doi:10.1063/1.880993.
  4. 1 2 Brown, Laurie M. (1986). "Hideki Yukawa and the meson theory". Physics Today. 39 (12): 55–62. Bibcode:1986PhT....39l..55B. doi:10.1063/1.881048.
  5. 1 2 Griffiths, David J. (2017). Introduction to Quantum Mechanics. Cambridge, United Kingdom: Cambridge University Press. p. 415. ISBN   978-1-107-17986-8.
  6. Müller, H.J.W. (1965). "Regge-Pole in der nichtrelativistischen Potentialstreuung". Annalen der Physik (in German). 470 (7–8): 395–411. Bibcode:1965AnP...470..395M. doi:10.1002/andp.19654700708.
  7. Müller, H.J.W.; Schilcher, K. (February 1968). "High-energy scattering for Yukawa potentials". Journal of Mathematical Physics. 9 (2): 255–259. doi:10.1063/1.1664576.
  8. 1 2 Müller-Kirsten, Harald J.W. (2012). Introduction to Quantum Mechanics: Schrödinger equation and path integral (2nd ed.). Singapore: World Scientific. ISBN   978-9814397735.
  9. Müller, H.J.W. (1965). "On the calculation of Regge trajectories in nonrelativistic potential scattering". Physica. 31 (5): 688–692. Bibcode:1965Phy....31..688M. doi:10.1016/0031-8914(65)90006-6.
  10. Langer, Rudolph E. (1937). "On the connection formulas and the solutions of the wave equation". Physical Review. 51 (8): 669–676. Bibcode:1937PhRv...51..669L. doi:10.1103/PhysRev.51.669.
  11. Boukema, J.I. (1964). "Calculation of Regge trajectories in potential theory by W.K.B., and variational techniques". Physica. 30 (7): 1320–1325. Bibcode:1964Phy....30.1320B. doi:10.1016/0031-8914(64)90084-9.

Sources