scholarly journals The collision between two electrons

It is well-known that the problem of the collision between two particles interacting according to the inverse square law is exactly soluble on the wave mechanics, and that the solution yields the same scattering laws as the classical theory. If, however, the two particles are identical, e.g. , two electrons or two α-particles, this is not necessarily the case; for the wave functions used must be antisymmetrical or symmetrical in the co-ordinates of the two particles; and this may affect the scattering laws. In this paper we shall discuss the collision between two particles possessing spin, such as electrons, and also between two particles without spin, such as α-particles. Assuming an inverse square law force between the particles, and neglecting the actual spin forces, we shall deduce from the symmetry properties of the wave functions a scatter­ing law differing considerably from the classical. We shall also mention the various methods by which the effect could be observed, and give some experi­mental evidence in its favour. The application of the exclusion principle to collision problems has been discussed by the author in a previous paper. Suppose we wish to describe the motion of two particles interacting in any field of force. We obtain a solution w (r 1 r 2 ) of the wave equation, where r 1 refers to the position of the first particle, and r 2 to that of the second. If we did not use antisymmetrical wave functions, we should argue that the probability that the first particle should be at r 1 and the second at r 2 would be | w (r 1 r 2 )| 2 , and therefore the probability that one particle should be at r 1 and the other at r 2 would be | w (r 1 r 2 )| 2 + | w (r 2 r 1 )|

1980 ◽  
Vol 35 (2) ◽  
pp. 252-253
Author(s):  
Fritz Bopp

Abstract A wave equation of a kind proposed by Pais in 1953 describes a particle with an infinite sequence of quantum states, which belong to the symmetrical representations (λ, 0) of the group SU 3. Particles composed of such single ones are connected with the whole set of representations (λ, μ) of SU 3. The wave equation is compatible with an exclusion principle. Assuming that only particles with zero triality occur, all quarks and quarklike particles are excluded. Neither coulours, nor bags are needed, as we do not need repulsive forces to exclude Li-atoms with symmetrical wave functions.


A method is developed for the solution of the wave equation for two electrons in the presence of two centres. The work of Lennard-Jones & Pople (1951) on the ground state of such a system is generalized so as to apply to all the excited states. Full advantage is taken of the symmetry properties of the wave functions, both in three-dimensional and six-dimensional space, to reduce the wave equation to a number of component parts, each of a particular symmetry type. This leads to sets of equations with characteristic symmetry properties appropriate to singlet states and triplet states, whether even or odd, positive or negative in the standard notation ( 1 ∑ - g ).


1. Gordon has proved that the scattering of particles by an inverse square field is the same on the wave mechanics as on the classical theory. Mott has however shown, from consideration of the symmetry of the wave functions, that the scattering is quite different from the classical when the scattering and scattered particles are identical. The scattering of alpha particles by helium is typical of the case of the collision of identical particles that have no pin and obey the Einstein-Bose statistics. For this case, the expression given by Mott for the number of alpha particles scattered between angles θ, and θ + d θ, in going a distance dr in a gas containing n atoms per unit volume, is N dθ dr = 4 πn N 0 e 4 / m 2 v 4 {cosec 4 θ + sec 4 θ + 2 cosec 2 θ sec 2 θ cos u } dθ dr , (1) where u = 8/137 c/v log cot θ, (2) and N 0 is the total number of incident particles, v, m and e , their initial velocity, mass and charge and c the velocity of light.


It has been shown by Mott on the basis of the wave mechanics, that in the case of collisions between identical particles the scattered particles should interfere with the projected particles travelling in the same direction. When α -particles are scattered in helium, if the scattered α -particles and projected helium nuclei of similar velocity are identical in all respects, there will be interference between the two streams of particles. For collisions in which the particles act upon each other with forces varying as the inverse square of the distance between them, the interference results in the scattering intensity varying above and below the classical value and rising to double the classical numbers at 45º. At small angles the scattering predicted by the quantum mechanics does not differ greatly from that given by the classical theory. An experiment carried out by Chadwick showed quite definitely that for sufficiently slow α -particles the amount of scattering at 45º was double that of the classical theory. For these α-particles of low velocity the results showed that the forces varied very little from Coulomb forces; hence it was evident that the discrepancy was due to the failure of the classical theory. The scattering of slow α -particles by helium has also been investigated by Blackett and Champion by means of an expansion chamber. The observed scattering was in good agreement with the wave mechanical scattering. These experiments verify the assumption upon which Mott s theory is based, namely, that it is impossible to distinguish between an α-particle and a nucleus of helium travelling at the same velocity. Thus the helium nucleus has no spin or vector quantity associated with it; its field of force is perfectly spherical.


The problem of the collision of two particles which act upon each other with forces varying as the inverse square of the distance between them has been solved exactly on the basis of the new quantum mechanics, and the solution is the same as that given by classical mechanics. In a recent paper, however, Mott has pointed out that this agreement between the predictions of classical mechanics and wave mechanics depends upon the dissimilarity of the colliding particles; if the particles are identical the scattering laws given by the wave mechanics will be very different from those of classical theory. Mott has considered the two types of collision between similar particles, (1) in which the particles possess spin, such as the collisions of electrons with electrons or protons with hydrogen nuclei, and (2) in which the particles have no spin, such as the collisions of α -particles with helium nuclei.


In the simplest cyclic system of π-electrons, cyclobutadiene, a non-empirical calculation has been made of the effects of configuration interaction within a complete basis of antisymmetric molecular orbital configurations. The molecular orbitals are made up from atomic wave functions and all the interelectron repulsion integrals which arise are included, although those of them which are three- and four-centre integrals are only known approximately. In this system configuration interaction is a large effect with a strongly differential action between states of different symmetry properties. Thus the 1 A 1g state is several electron-volts lower than the lowest configuration of that symmetry, whereas for 1 B 1g the comparable figure is about one-tenth of an electron-volt. The other two states examined, 1 B 2g and 3 A 2g are affected by intermediate amounts. The result is a drastic change in the energy-level scheme compared with that based on configuration wave functions. Neither the valence-bond theory nor the molecular orbital theory (in which the four states have the same energy) gives a satisfactory account of the energy levels according to these results. One conclusion from the valence-bond theory which is, however, confirmed, is the somewhat unexpected one that the non-totally symmetrical 1 B 2g state is more stable than the totally symmetrical 1 A 1g . On the other hand, it is clear that the valence-bond theory, with the usual value for its exchange integral, grossly exaggerates the resonance splitting of the states, giving separations between them several times too great. Thus the valence-bond theory leads to large values of the resonance energy (larger, per π-electron, than in benzene) and so associates with the molecule a considerable π-electron stabilization. This expectation has no support in the present more detailed and non-empirical calculations.


1. In a previous paper, the author has shown that the anomalous scattering of α-particles in helium can be accounted for, if one postulates the validity of the wave mechanics to describe the phenomenon, and assumes a simple spherically symmetrical form for the field between the two colliding particles. The solutions obtained were exact, no approximate methods being used, but only the scattering at large angles, 34° and 45°, was considered. The purpose of this paper is threefold. Firstly, these results are extended to the case of the scattering of α-particles by hydrogen, and good agreement with experiment is obtained. Secondly, it is shown that the scattering at small angles both in hydrogen and in helium can be explained by the same field as is used to explain the scattering at large angles. It is therefore no longer necessary to assume a “plate-like” form for the α-particle, in order to explain the scattering at small angles, as formerly according to the classical mechanics. Thirdly, a discussion is given of the extent to which the explanation here advanced for the experimental results is dependent on the particular form assumed for the potential energy of the one particle in the field of the other. It is found that, provided we assume that the potential energy, V( r ), is Coulombian for distances greater than about 5 X 10 -13 cm., then whatever the form of the potential for smaller distances , the ratio, R, of the scattering to that to be expected from a purely Coulombian field is given by the formula R = | e - i / ak . logcos 2 א + iak cos 2 א (e 2iK 0 — 1)| 2 , where the scattering is in hydrogen, א is angle of scattering, and ak = hv /4πє 2 . A similar formula is found for the scattering in helium. The formula contains only one parameter, K 0 , which depends on the velocity of the incident α-particles, v , and also on the field assumed, but is independent of the angle of scattering, א. Thus we may determine K 0 from the observed scattering for a given value of v and of א, and deduce the scattering at other angles for the same value of the velocity. The agreement obtained with experiment is then quite independent of any special choice of a potential energy function. In order to calculate K 0 and its variation with v , we should have to assume a particular form for V( r ) Conversely, from the values of K 0 determined by the experimental values of R, we may calculate V( r ) as in the previous paper for the case of two α-particles. 2. The Scattering Formula for Hydrogen .—In the experiments on the scattering by hydrogen, it is found convenient to count the scintillations due to the hydrogen nuclei projected forward by the collision rather than those due to the deflected α-particles themselves, and therefore in this portion of the paper we shall denote by It the ratio of the number of protons found experimentally to be projected in a given direction, to the number predicted by the Rutherford law. The fact that R refers to the particles struck on from rest makes a considerable simplification in the collision relations, for if we consider the collision of two particles of mass m 1 and m 2 , of which m 1 is initially at rest, and if we denote by א the angle made by the final direction of motion of this particle with the incident direction of m 2 , and by θ the angle through which either particle is deflected, measured in a frame of reference moving with the velocity of the centre of gravity of the two particles, then א = ½ (π - 0), (1) independently of the ratio m 1 / m 2 . The relation between the angle θ and the angle ϕ, through which the incident particle m 2 is deflected, is on the other hand dependent on the relative masses of the particles.


The study of the scattering of α-particles is now a familiar method of finding the fields of force of light nuclei, because the inverse square law of repulsion breaks down at distances to which α-particles are able to approach them. It has been shown how wave mechanics may be applied to the interpretation of experimental data in terms of an interaction potential between the nuclei which deviates from the Coulomb value at separations of the order 3 × 10 -13 cm. This study is valuable, for with knowledge about the simplest particles one may more easily approach the difficult problem of the structure and stability of the heavier nuclei. The previous experimental work on the scattering of α-particles in helium by Rutherford and Chadwick (1927), and by Chadwick (1930), and in hydrogen by Chadwick and Bieler (1921) cannot be regarded as sufficiently precise, and for deuterium we have only the preliminary experiments of Rutherford and Kempton (1934), and of Pollard and Margenau (1935). Wright’s (1932) results for helium are more definite, but even they do not complete the desired information for the slower α-particles. For hydrogen and deuterium no means were used to obtain accurate absolute values.


1928 ◽  
Vol 24 (3) ◽  
pp. 445-446 ◽  
Author(s):  
H. D. Ursell

A simple explanation of Pauli's principle was first given with the wave mechanics. Its interpretation in the new theory was that the wave functions of Schrödinger were antisymmetrical in all the electrons concerned. Thus when the interactions of the electrons may be neglected, the wave function (for a system of n electrons) can never be of the formin nature, but only of the formHence σ, τ,…ω must be all distinct.


Author(s):  
Frank S. Levin

Chapter 7 illustrates the results obtained by applying the Schrödinger equation to a simple pedagogical quantum system, the particle in a one-dimensional box. The wave functions are seen to be sine waves; their wavelengths are evaluated and used to calculate the quantized energies via the de Broglie relation. An energy-level diagram of some of the energies is constructed; on it are illustrations of the corresponding wave functions and probability distributions. The wave functions are seen to be either symmetric or antisymmetric about the midpoint of the line representing the box, thereby providing a lead-in to the later exploration of certain symmetry properties of multi-electron atoms. It is next pointed out that the Schrödinger equation for this system is identical to Newton’s equation describing the vibrations of a stretched musical string. The different meaning of the two solutions is discussed, as is the concept and structure of linear superpositions of them.


Sign in / Sign up

Export Citation Format

Share Document