Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Quantum Mechanics of One- and Two-Electron Atoms
Quantum Mechanics of One- and Two-Electron Atoms
Quantum Mechanics of One- and Two-Electron Atoms
Ebook837 pages15 hours

Quantum Mechanics of One- and Two-Electron Atoms

Rating: 0 out of 5 stars

()

Read preview

About this ebook

This classic of modern physics includes a vast array of approximation methods, mathematical tricks, and physical pictures that are also useful in the application of quantum mechanics to other fields. Students and professionals will find it an essential reference for calculations pertaining to hydrogen-like and helium-like atoms and their comparison with experimental results.
In-depth explorations of the Dirac theory of the electron and of radiative effects include brief accounts of relevant experiments. The specific application of general field-theoretic results to atomic systems also receives a thorough examination. Author Hans A. Bethe (1906–2005), Professor of Physics at Cornell University, won the Nobel Prize in Physics in 1967. Co-author Edwin E. Salpeter is James Gilbert White Distinguished Professor of the Physical Sciences at Cornell University.
LanguageEnglish
Release dateApr 26, 2013
ISBN9780486318288
Quantum Mechanics of One- and Two-Electron Atoms

Related to Quantum Mechanics of One- and Two-Electron Atoms

Titles in the series (100)

View More

Related ebooks

Physics For You

View More

Related articles

Reviews for Quantum Mechanics of One- and Two-Electron Atoms

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Quantum Mechanics of One- and Two-Electron Atoms - Hans A. Bethe

    Introduction. Units.

    α) One of the simplest, and most completely treated, fields of application of quantum mechanics is the theory of atoms with one or two electrons. For hydrogen and the analogous ions He+, Li++, etc., the calculations can be performed exactly, both in SCHRÖDINGER’S nonrelativistic wave mechanics and in DIRAC’S relativistic theory of the electron. More specifically, the calculations are exact for a single electron in a fixed COULOMB potential. Hydrogen-like atoms thus furnish an excellent way of testing the validity of quantum mechanics. For such atoms the correction terms due to the motion and structure of atomic nuclei and due to quantum electrodynamic effects are small and can be calculated with high accuracy. Since the energy levels of hydrogen and similar atoms can be investigated experimentally to an astounding degree of accuracy, some accurate tests of the validity of quantum electrodynamics are also possible. Finally, the theory of such atoms in an external electric or magnetic field has also been developed in detail and compared with experiment.

    For atoms and ions with two electrons, such as H−, He, Li+, etc., exact analytic calculations are not possible at the present time. But these atoms are still simple enough, so that various approximation methods can be used to carry out calculations to a high degree of accuracy. In fact, for the ground state of such atoms the accuracy of the theoretical calculations is of the same order of magnitude as that of spectroscopic measurements. Thus, helium-like atoms not only give further confirmation of the general validity of quantum mechanics, but provide an excellent testing ground for the various approximation methods which are commonly used in quantum mechanics. Finally, the relativistic effects for two- electron atoms (and especially for positronium) provide a simple field of application for present-day theories of the interaction between electrons.

    We shall largely base our treatment on the standard wavemechanical form of quantum mechanics in the nonrelativistic theory and on the DIRAC theory for the relativistic treatment of electrons (particles of spin ). The fundamental principles of these theories are described in detail elsewhere, see for instance the works¹ by DIRAC [1], KRAMERS [2], PAULI [3] or SCHIFF [4], Occasionally we shall use the more general methods of matrix manipulation, described, for instance, by CONDON and SHORTLEY [5], We shall not treat quantum electrodynamics in any detail, but will refer to results obtained in this theory (see for instance, HEITLER [6]).

    β) Units. To avoid carrying too many numerical factors, we shall in general use atomic units (a.u.), introduced by HARTREE². These units are built up of various combinations of the charge e and mass m of the electron and of PLANCK’S constant h. In general we shall find it useful to introduce, instead of h, the rationalized PLANCK’S constant,

    We shall also use SOMMERFELD’S dimensionless fine structure constant

    where c is the velocity of light, c = 299 793 km/sec. In our atomic units, ħ is also the unit of action and of angular momentum.

    The atomic units are:

    1. Unit of charge = e = charge of the electron = 4.8029 × 10−10 e.s.u. = 1.6021 × 10−20 e.m.u.

    2. Unit of mass = m = mass of the electron = 9.1085 × 10−28 gm.

    3. Unit of length = a = "radius of first BOHR orbit" (innermost circular orbit for H in the old quantum theory) = ħ²/me² = 5.2917 × 10−9 cm.

    4. Unit of velocity = v0 = electron velocity in the first BOHR orbit = e²/ħ = αc = 2.1877 ×10⁸ cm/sec.

    5. Unit of momentum = p0 = electron momentum in the first BOHR orbit = me²/ħ = mv0 = 1.9926 × 10−19 gm cm/sec.

    6. Unit of energy = e²/a = me⁴/ħ² = = twice the ionization potential of hydrogen (for nuclear mass ∞) = 4.3590 × 10−11 erg.

    7. Unit of time = a/v0 = ħ³/me⁴ = 2.4189 × 10−17 sec.

    8. Unit of frequency = v0/a = me⁴/ħ³ = Ry = 4.1341 × 10¹⁶ sec−1 (Ry = RYDBERG frequency).

    9. Unit of electric potential = e/a = me³/ħ² = 0.09076 e.s.u. = 27.210 Volt.

    10. Unit of electric field strength = e/a² = m²e⁵/ ⁴ = 5.142 × 10⁹ Volt/cm.

    We have defined the RYDBERG for infinite mass Ry as a frequency, namely 1/4π atomic frequency units. We shall sometimes express energy in terms of RYDBERG units, that is in units of (h Ry). This quantity is one half an atomic energy unit and numbers expressed in terms of it will carry the symbol Ry after them. In many experimental results, one obtains instead of an energy E, the equivalent wave number (E/hc), expressed in cm−1, or the equivalent frequency (E/h), expressed in Mc/sec. The wave number corresponding to one RYDBERG unit of energy is³

    The RYDBERG frequency itself, equal to 3.28 985 × 10⁹ Mc/sec.

    One RYDBERG unit of energy, expressed in terms of electron volts, is 13.6050 eV.

    γ) Basis for numerical values used. The numerical values quoted above for the atomic units expressed in ordinary (C.G.S.) units were taken from DUMOND and COHEN’S⁴ Tables of the Physical Constants. DUMOND and COHEN’S best numerical values of the physical constants are based on a large number of experiments, which measure different combinations of the fundamental constants. More experiments than unknown constants were available and the best values were derived by a method-of-least-squares analysis. Some of the most accurate experiments, used in this analysis, require for their interpretation a knowledge (and the validity) of the quantum theory of atomic energy levels. This theory, to be developed in the following chapters, is required in particular (a) to derive the RYDBERG, R∞ from the spectroscopic measurement of wave numbers of the BALMER lines for hydrogen, (b) to derive the fine structure constant from microwave measurements of the fine structure (or hyperfine structure) splitting of energy levels of hydrogen.

    To be able to discuss experimental tests for the validity of atomic theory, it is useful to consider the best values of the atomic constants which can be derived from experiments without any recourse to atomic theory. Besides the accurately known value for the velocity of light, c, we need numerical values for h, e and m. Enough experiments for such a determination are now available: In particular PLANK’S constant h can be derived from measurements of the short wave-length limit of the continuous X-ray spectrum without detailed knowledge of quantum theory, using only the postulate E = hv, relating energy E and frequency v. An outline of some of the other experiments is as follows: X-ray diffraction experiments on crystals and gratings give the lattice constants of simple crystals in centimeters and hence AVOGADRO’S number N. The measurement of the Faraday then gives e. Mass-spectroscopy (and N) gives Mp, the proton mass. Experiments comparing (indirectly) the cyclotron frequencies of the electron and proton finally give m/Mp, and hence m.

    The numerical values obtained from these experiments, which do not involve atomic theory for h, e and m (and hence for R∞ and α) are, of course, less accurate than DUMOND and COHEN’S best values (using all experiments) which were quoted above. These non-atomic values⁵ are

    These non-atomic values agree with the more accurate ones, quoted above, to within (approximately) the experimental error (the discrepancy is less than 1.5 times the standard deviation in all cases)⁶.

    ¹ Numbers in italics, e.g., [1], refer to a bibliography at the end of this article.

    ² D. R. HARTREE: Proc. Cambridge Phil. Soc. 24, 89 (1928).

    ³ This quantity R∞ is often referred to simply as the RYDBERG for infinite nuclear mass in the literature.

    ⁴ J. W. M. DUMOND and E. R. COHEN: Rev. Mod. Phys. 25, 691 (1953); also first article in this volume.

    ⁵ J. W. DUMOND and E. R. COHEN: In this volume and private communication.

    ⁶ Numerical values, found in older theoretical references (including ref. [10] of our bibliography), which are based on the atomic constants, should be treated with great caution: The older values of some atomic constants were in error by much more than the statistical errors in the older experiments (due to systematic errors not expected at the time).

    I. The hydrogen atom without external fields.

    a) Nonrelativistic theory.

    1. Separation of SCHRÖDINGER’S equation in spherical polar coordinates. Angularly dependent eigenfunctions and the angular momentum matrix. SCHRÖDINGER’S equation¹ in c.g.s. units for an electron in the field of a nucleus of charge Ze and of infinite mass is

    and in HARTREE’S atomic units (see Introductory Remarks) is

    This equation can be separated in spherical polar coordinates. We choose the nucleus as origin of a polar coordinate system with some arbitrary z direction as polar axis. Let be the coordinates of the electron. The LAPLACIAN operator Δ is eriven by

    We try as solution

    and obtain

    The left side is a function of r only, the right side a function of and only. Therefore, λ is a constant. The equation on the right in (1.4), namely

    can be solved only if

    In this event there are the following 2l + 1 solutions:

    Ylm is a spherical harmonic, the unnormalized associated LEGENDRE function (cf. Appendix).

    The first few spherical harmonics are given explicitly:

    The normalized associated LEGENDRE functions for l = 1 to 3 are graphed in Fig. 1 (i.e., the parts of the spherical harmonics depending on geographic latitude ). As one can see, the associated LEGENDRE functions Plm have l − |m| zeros between the poles of the sphere. Thus, the spherical harmonic Ylm has l − |m| parallels of latitude as nodal lines. Furthermore, the real part of the spherical harmonic

    has |m| meridians as nodal lines, resulting in a total of l nodal lines.

    The eigenfunctions

    are closely related to the angular momentum of the atom—it turns out that both the total angular momentum and the angular momentum about the z-axis are diagonal matrices. They are quantized; i.e., if the appropriate operators are applied to the eigenfunction u,

    Fig. 1 a—c. The normalized associated LEGENDRE function plotted against (in degrees).

    the result is the eigenfunction u multiplied by a constant. The operator belonging to the angular momentum about the z-axis is given by²

    Operating on u, one obtains

    m is a measure of the angular momentum in the z direction and is called the magnetic quantum number³.

    The operator representing the total angular momentum is given by

    resulting from an elementary transformation. In view of (1.2) through (1.6) it follows that

    We could have saved ourselves the trouble of calculating the angularly dependent eigenfunctions explicitly, by making use of some general theorems about the angular momentum. These theorems state: 1. The components and the absolute value of the total angular momentum of an electron in any central force field are constants of the motion. 2. The eigenvalues belonging to the square of the total angular momentum are equal to l(l + 1), where l is an integer. 3. The eigenvalues of the components in a fixed direction are equal to m, where m can assume all integral values from −l to +l. 4. A quantum state is defined by specifying the magnitude and one of the components of the total angular momentum, i.e., by specifying l and m. Thus, by means of the generally valid formula (1.11) one can arrive at once at the differential equation for the radially dependent eigenfunction

    which is identical with the left side of (1.4).

    Before solving this equation, we shall calculate the matrix elements of the angular momenta kx and ky about the directions perendicular to the z-axis. By definition

    Thus, in view of formulas (A. 25) and (A. 26) of our appendix,

    and the matrix elements of the components of the angular monentum become

    The matrix elements which correspond to transitions between states of different orbital quantum numbers or of different radially dependent eigenfunctions are zero. As is well known, the formulas (1.14), except for an indeterminancy in the sign of the square root, can be obtained from the general theory (ref. [3] of the bibliography).

    We discuss finally the parity of our wave functions: For a one-electron atom the parity operation consists simply in the replacement of the position vector r of the electron by −r. In spherical polar coordinates this means replacing by It follows from the properties of spherical harmonics, given in our appendix, that

    For even l, then, our wave functions are unaffected by the parity operation and we say the parity is even. For odd l, the wave functions change sign under the parity operation and the parity is odd.

    2. Derivation of BALMER’S formula ⁴. We shall now treat the differential equation satisfied by the radially dependent part of the eigenfunction

    The last term in the above equation corresponds classically to the centrifugal force of the electron and increases with increasing angular momentum.

    Let us begin by assuming that the energy E is negative. Since the potential energy vanishes at infinity, a negative energy corresponds to a bound electron which possesses a positive kinetic energy only by virtue of the nuclear attraction.

    First, we study the asymptotic behavior by neglecting the terms in r which are of lower power compared to the ones that are kept, then

    If R(∞) is to remain finite, we must select the minus sign. We introduce the notation

    and extend the asymptotic solution (2.2) to a solution valid for all r by putting

    where f(r) is a function which varies slowly at infinity. Substitution of (2.4) into (2.1) results in a differential equation in f

    Let us expand f in a power series,

    Substitution of (2.6) into (2.5) results in

    The coefficient of each power of r must vanish in the above equation. Setting the coefficient or −² (the term of lowest power in r) equal to zero, one obtains an equation which determines λ:

    from which

    The condition that f remain finite at r = 0 forces one to the choice λ = l. Setting the coefficient of + v − ² equal to zero one obtains the recursion formula

    Next, we require that f be a polynomial in r, i.e., that the power series terminate, an l − 1rn − ¹, say, being the last term. Then

    This condition will be fulfilled if

    If the above requirement were not made, then for large values of v:

    where c is a constant; f(r) would behave as e² er for large r and R = eεr f(r) would increase as e+εr. Therefore, the breaking off of the series is necessary to assure that the eigenfunction be bounded at infinity.

    (2.11) is the well-known BALMER formula for the discrete energy levels of hydrogen (Z = 1) and for the ions having a single electron, such as He+, Li++ etc. n is the principal quantum number. The energy does not depend on the other two quantum numbers m and l. The independence of the magnetic quantum number m has its origin in the fact that all directions in space enter on equal terms. This holds for all atoms in the absence of an external field and is called directional degeneracy. On the other hand, the fact that the energy is independent of the quantum number l is a special property of the hydrogen atom which must be attributed to the presence of the exact COULOMB potential Z/r. This degeneracy has the consequence that the hydrogen atom is particularly strongly influenced by external fields (first-order STARK effect instead of the usual second-order effect⁵).

    The energy formula (2.11) has been verified to an extraordinary degree by spectroscopic measurements. The spectral line emitted in a transition of the atom from quantum state n to state n′ has a frequency

    in atomic units⁶, and is equal to

    Fig. 2. The energy level scheme of hydrogen (both the energies in electron volts and the corresponding wave numbers in cm−1 are given).

    where the RYDBERG frequency Ry was defined in the Introduction. The term scheme of hydrogen is shown in Fig. 2. The LYMAN, BALMER, PASCHEN, and BRACKET series, which have lowest levels corresponding respectively to the quantum numbers n′ = 1, 2, 3, 4, are shown in Fig. 3. Figs. 2 and 3 are taken from

    Fig. 3. The spectral lines of the hydrogen spectrum (the height of a line is proportional to its frequency, the number written next to a line is its wave length in ANGSTROM units).

    GROTRIAN’S book Graphische Darstellung der Spektren von Atomen und Ionen. It is apparent that the Jines in the LYMAN series are more densely spaced than the lines in the higher series.

    Usually, instead of its frequency v, the wave number (inverse of the wavelength) of a spectral line is measured, which is simply (v/c), where c is the velocity of light, 299793 km/sec. The wave numbers of many lines of the spectrum of the hydrogen atom can be measured to an accuracy of better than one part in a million. The Hα lines in the hydrogen and deuterium spectrum, in particular, have been measured and analyzed very carefully. The exact theoretical expression for the frequency of the Hα line in an actual atom differs slightly from that obtained from (2.13). This difference is due to correction terms allowing for the finite mass of the nucleus, the relativistic fine structure (and hyperfine structure) splitting of the energy levels and quantum electrodynamic level shifts (all to be discussed later). These correction terms are small and well known. COHEN⁷ has analyzed the experimental information on the Hα lines in H and D and applied the appropriate correction terms to (2.13) obtain aft extremely accurate value or the RYDBERG constant for infinite nuclear mass. The present value is⁸

    Using this value for Ry, (2.14), obtained from the Hα line, the measured wave numbers of other spectral lines in hydrogen can be compared with the theoretical values, (2.13). After applying corrections for fine structure, etc., the agreement for a large number of spectral lines is excellent, to better than one part in a million for the Hβ line, for instance. As pointed out in the Introduction, the atomic constants e, m and h, entering into the definition for Ry could also be obtained from experiments which do not require atomic theory for their analysis. Such a non-spectroscopic value for Ry would only be accurate to about one part in three thousand, but agrees with (2.14) to this accuracy. Many spectral lines have also been measured for hydrogen-like ions of higher nuclear charge Z, up to seven-times ionized oxygen (Z = 8). The agreement is again excellent⁹.

    3. The radial eigenfunctions of the discrete spectrum¹⁰. We shall now deal with the radial eigenfunctions in greater detail. Combining (2.9) with (2.10)

    and thus

    11

    where by definition

    and F is the (confluent) hypergeometric function¹²

    c is a constant.

    The radial eigenfunctions can also be expressed in terms of the associated LAGUERRE functions, which are defined by means of the relations¹³

    This may be seen by carrying out the indicated differentiations; then

    On comparison with (3.2) one finds that

    Naturally, one can also prove directly that (3.8) satisfies the differential Eq. (2.1).

    Next, we require that the eigenfunction be normalized according to the rule

    i.e.,

    R is now the normalized eigenfunction.

    α) Evaluation of integrals over LAGUERRE functions¹⁴. Instead of evaluating (3.10), we evaluate the more general integral:

    For this purpose we substitute the representation (3.5) for one of the LAGUERRE functions:

    Integrating by parts μ times with respect to one obtains

    We must now distinguish between two cases: left with the evaluation of the integral. We proceed by replacing the LAGUERRE function with its power series expansion (3.7) and then carry out the μ fold differentiation :

    Integrating by parts λ more times results in

    the factor λe− assuring the vanishing of the integrated parts. A final integration with respect to 9 gives

    The summation over γ can be accomplished by elementary means and gives

    Introducing the notation β = α λ + μ + σ one finally gets

    2. For negative σ the reverse is true. The integral vanishes, as is evident from (3.13), and the integrated terms make non-zero contributions at the lower limits (the upper limits still go to zero exponentially). By means of (3.7) one obtains

    Substitution into (3−12) gives

    The summation over β can be accomplished by elementary means and gives

    Putting − (σ + 1) = s and s α = γ one finally gets

    β) Discussion of the normalized eigenfunctions. The normalization integral (3.10) can now be evaluated by setting λ = n + l, μ = 2l and σ = 1 in formula (3 13):

    Thus, the normalized eigenfunction [cf. (3.2), (3.8), (3.10) becomes:

    We write down explicit expressions for the first few radial eigenfunctions of hydrogen (Z = 1):

    A few of the eigenfunctions are graphed as functions of in Fig. 4. Fig. 5 represents the charge distributions corresponding to these eigenfunctions. We have plotted as a function of r, i.e. the probability that the electron be found in the spherical shell between r and r + dr. It is apparent that the maximum charge density recedes from the nucleus with increasing principal quantum number n. For eigenfunctions which have the same n, the one with lowest l has the largest amplitude in the neighborhood of the nucleus. The radial eigenfunction has n − l − 1 zeros, and thus the complete eigenfunction

    has n − 1 nodal surfaces of which n l − 1 are concentric spheres (r = const), − m are cones ( = const) with common apex at the origin, and m are planes

    Fig. 4. The radial eigenfunctions (according to PAULING). The abscissa is = 2r/n, the ordinate is nRnl(r). The symbols K, etc. refer to the corresponding X-ray levels (in X-ray terminology the symbols L11, L21, L22, are usually replaced by LI, LII, LIII).

    Fig. 5 a—c. The charge distribution in the first few states of the hydrogen atom. The abscissa is r in atomic units, the ordinate is the charge density The curves are labelled with the corresponding nl values (s-states in Fig. 5a, p-states in Fig. 5b and d- and f-states in Fig. 5c).

    through the origin (φ = const)¹⁵. A particularly beautiful pictorial representation of the charge distribution in the various states of hydrogen is afforded by the silhouettes of WHITE¹⁶.

    γ) Mean values of powers of r. The differences between the various radial eigenfunctions become particularly apparent when we form the mean value of r, the distance between electron and nucleus, raised to various powers. The mean value rv is given by

    in which is the quantity given in (3.13) and (3.14). Assigning particular values to the indices results in the explicit formulas

    The mean values for the first few states of hydrogen are given in Table 1.

    Table 1. Numerical values for the expectation values of rv.

    The mean values of r raised to positive powers v are determined essentially by the principal quantum number n, while for the mean values of r raised to negative powers v(for v −1) the orbital quantum number l becomes decisive. This is readily explained. For positive powers the important contributions to the integral come from the large values of r for which the eigenfunction behaves like for small r, important for negative powers, the eigenfunction behaves like l. In terms of a picture this means that the probability of finding the electron at a large distance from the nucleus is essentially the same for both the circular (l = n − 1) and the eccentric elliptical (small l) BOHR orbits. For fixed n the electron will be found more frequently in the immediate vicinity of the nucleus if its quantum state corresponds to an eccentric orbit than if its quantum state corresponds to a circular orbit. (It is even possible for to diverge, as indeed happens if v < − 2l − 2; the mean value then becomes meaningless.)

    The quantum mechanical analogue to BOHR orbits of large eccentricity (small l) is a large mean square deviation of the nucleus-electron separation:

    For example

    As can be seen from (3.24) the expression for the mean value of r−1 is particularly simple. By means of it we can readily verify the well-known virial theorem¹⁷ which states that the mean value of the potential energy is

    equal to twice the total energy. Thus,

    δ) Behavior of the eigenfunctions for large principal quantum number. WENTZEL- KRAMERS-BRILLOUIN (WKB) method. Aside from normalization, all eigenfunctions of a fixed orbital quantum number l and different principal quantum number n behave alike near the nucleus, provided n l. This may be seen by neglecting l relative to n in formula (3.17), thus,

    or it can be seen, even more simply, directly from SCHRÖDINGER’S Eq. (2.1). If n is very large and r is of order 1 (more precisely, if r n²/Z), the energy 1/2n² can obviously be neglected compared to 1/r and l(l + 1)/r². Then (2.1) goes over into

    Since the differential Eq. (3.31) no longer contains n, its solution will also be independent of n. Its solution is

    where J2l+1 is the BESSEL function¹⁸ of order 2l + 1 and c is a constant. If the power series expansion for J is used in (3.32), then the expansion (3.30) for R is again obtained. Using the asymptotic formula for the BESSEL function one obtains a solution valid for large r

    One must be sure, however, that r n²/Z (but 8Z r 1).

    A useful approximate formula for the radial eigenfunction may be obtained by means of the WKB method when r becomes comparable to n²/Z (cf. Sect. 53). First, we must get rid of the first derivative in Eq. (2.1). For this purpose we introduce v = Rr in place of R. v satisfies the equation

    The coefficient of v represents the kinetic energy of the electron and is positive for r1 < r < r2 where

    r1 and r2 are respectively the perihelion and aphelion of the classical electron orbit. In the region of the classical orbit r1 < r < r2 the eigenfunction may be represented according to (53.3) to a good approximation by¹⁹:

    Although (3.36) looks rather complicated it turns out to be a useful formula in practice. For (3.36) becomes

    The radial eigenfunction R = v/r thus becomes identical with (3.33), as must happen²⁰, and the constants are related through

    v decays exponentially outside of the region of the classical orbit [see (53.4)]. Thus, if we wish to evaluate the normalizing integral

    we only need to consider the region r1 < r < r2. In addition, we can make use of the fact that the cosine in (3.36) is a rapidly varying function compared to the other factor, and replace cos² by its mean value . Then

    The eigenfunction is thus proportional to and, furthermore, this is the only factor through which u depends on n. For large n the normalization (3.39) is, of course, identical with the one previously derived. If the value for c is substituted into (3.32) and the series expansion for the BESSEL function (as given, for example, in JAHNKE-EMDE) is used, then exactly (3.30) is obtained.

    ε) Generating function for LAGUERRE functions²¹. Occasionally a representation of the LAGUERRE polynomials different from (3.5) is preferable, particularly when we are dealing with a calculation of transition probabilities²². It turns out that the LAGUERRE polynomials can be defined by means of the generating function

    Proof: Differentiating (3.40) with respect to t and setting the coefficient of th eaual to zero gives

    Differentiating (3.40) with respect to x, one obtains

    By means of a short calculation the differential equation

    follows from (3.41), (3.42). If we set k = n + l and differentiate 2l + 1 times with respect to x, we arrive at the equation

    Putting the differential equation (2.5) follows without difficulty in view of (2.10).

    (3.40) agrees with (3.5) also with respect to normalization as can be shown by calculating the constant term in Lk, i.e., the term which is free of x, thus

    which is in agreement with (3.6). The associated LAGUERRE functions can also be represented by a generating function. Differentiating (3.40) r times with respect to x we obtain

    We finally quote the value at the origin of the squares of the normalized radial and total wave functions for l = 0 and any principal quantum number n. From (1.8) and (3.2) to (3.4) we have

    4. The eigenfunctions of the continuous spectrum²³, α) We shall now treat the case E > 0. The quantity ε, defined in (2.3), is now purely imaginary

    Fig. 6. Integration contour for the continuous eigenfunctions. The path a encloses both branch points ξ = + (x = 0) and ξ = − (x = − ) of the integrand. Integration along the path b gives (for large r) the asymptotic incoming wave Rt, the path c gives the outgoing wave R(1)

    In every other respect there is no change in the derivation of Sect. 2 up to the recursion formula (2.9). The ratio av/av − 1 becomes complex, and it is no longer possible to terminate the series (2.6) through a particular choice of ε. On the other hand, there is no need for that now because both e+εr and eεr remain finite for r = ∞. Thus, there is a solution for every positive E, and a continuous spectrum of positive eigenvalues adjoins the discrete levels of negative energy.

    The eigenfunctions belonging to the states of the continuous spectrum are, in accordance with (3.2) and (3.4), expressible in terms of hypergeometric functions since the recursion formula (2.9) is still valid. However,

    is imaginary.

    Accordingly, the definition (3.5) of the LAGUERRE polynomials has no longer any apparent meaning. However, it is possible, using CAUCHY’S theorem,

    to obtain a representation of the LAGUERRE polynomials,

    which is valid for all λ. In the second of the above representations the variable of integration z has been replaced by x + . The path of integration consists of the simple loop a (Fig. 6) around the two branch points x = − and x = 0 of the integrand, μ-fold differentiation of (4.4) results in the integral representation of the associated LAGUERRE functions (μ is integral and positive):

    Finally, the radial eigenfunction is obtained by setting λ = n + l, μ = 2l + 1, multiplying by and including all constant factors in the constant c:

    The representation (4.7) is obtained from (4.6) by substituting the path of integration going around the branch points in the positive sense. From (4.7) it is at once apparent that the radial eigenfunctions will be real if and n are pure imaginaries.

    Expanding (4.6) in powers of and carrying out the integrations by means of CAUCHY’S theorem, the representation (3.2) for the radial eigenfunction is again obtained (within a numerical factor):

    The series for the hypergeometric function F [cf. (3.4)] converges for all , but the convergence is very slow for large . Therefore, it is necessary to look for an asymptotic series in descending powers of , which will be useful for large . For this purpose we deform path a (Fig. 6) into two loops b and c coming from infinity and each going around one of the branch points in the positive sense. Along path b we expand (x + )n l − ¹ in decreasing powers of :

    This expansion diverges for the remote portions (|x| > | |) of the path of integration; the resulting asymptotic expansion for Rnl is, therefore, only semi-convergent. For the actual integrations we refer the reader to the work of SOMMERFELD and SCHUR and simply state here the results: The contribution of path b to the asymptotic representation of the eigenfunction is

    where

    is a hypergeometric function²⁴. Along path c it is convenient to replace x by z [cf. (4.4)]. The contribution to the asymptotic representation of the eigenfunction turns out to be exactly the complex conjugate of .

    Collecting the contributions and substituting the values of and n from (3.3) (4.1) and (4.2), the asymptotic expression for the wave-function is obtained:

    where

    is the complex phase of the Г-function. Thus, asymptotically, the eigenfunctions go over into spherical waves.

    We must now normalize the eigenfunction R. The well-known rule for the normalization of eigenfunctions belonging to the continuous spectrum is

    In the above, T is any function of the wave number k, e.g. the energy W = k² or k itself. Δ T is a small interval. If condition (4.11) is fulfilled and the eigenfunctions belonging to the discrete spectrum are normalized in the usual fashion,

    then an arbitrary function of the space coordinates can be expanded in terms of our eigenfunctions as follows:

    and the coefficients in the expansion are given by

    The eigenfunctions RTl are said to be normalized in the T-scale. The eigenfunctions normalized in the k-scale and k-scale respectively are related through

    as follows directly from (4.11).

    We calculate the normalizing factor in the k-scale by putting

    in accordance with (4.10); b is the normalizing constant to be determined, and δl is independent of r. If we neglect quantities of order 1/kr and Δk/k we obtain

    Substitution of (4.15) and (4.16) into (4.11), setting T = k, and replacing the rapidly oscillating cos² by its mean value results in

    Thus, in the k-scale

    Normalizing in the energy-scale we get, correspondingly,

    Comparing the normalized eigenfunction (4.18) with the asymptotic representation (4.10) for the unnormalized eigenfunction we find that

    At this point we can find both the integral and, for small r, the series representations of the normalized eigenfunction, the former by going back to (4.6) and (4.7) and the latter by substituting (4.20) into (4.8). In doing this it is practical to write the Γ-function in terms of elementary functions by means of the well known recursion relation

    and the formula

    then²⁵

    We thus obtain our final integral (4.22) and series (4.23) representations:

    Fig., 7 shows the eigenfunctions belonging to the continuous spectrum for E = 0, 0.25 and 1 (in Ry units).

    Fig. 7. The continuous eigenfunction of hydrogen for s-states (l = 0) for three different energies E (in Ry units). Abscissa is r in atomic units, ordinate is rRE0. Note that the wave length and amplitude increase rapidly with increasing r for E = 0, but are almost constant for E = 1.

    β) The irregular functions. In this section we have discussed wave functions for the continuous spectrum in the form of a radial wave function R(r) multiplied by a spherical harmonic Ylm. For any central potential (function of the radial distance r only) the SCHRÖDINGER equation can be separated in spherical polar coordinates and wave functions of the form Rl(r) Ylm exist. We have only calculated the radial wave functions for the special case of the COULOMB potential Zr−1. Further, we have only treated radial wave functions which are finite at the origin, the regular COULOMB functions. There exists another set of solutions forthe radial wave functions which diverge at the origin. These irregular COULOMB functions have an asymptotic form representing spherical standing waves, similar to that of the regular functions (4.10), but with cosine replaced by sine. These irregular functions do not arise in physical problems involving only pure COULOMB potentials, since physical wave functions must remain finite at the origin. They are of use, however, in problems involving central potentials which approximate a COULOMB potential at large distances, but deviate from it for small distances. These irregular functions occur²⁶, for instance, in nuclear problems but not in atomic theoiy and we shall not consider them further.

    5. Motion of the nucleus. So far we have pretended (see Sect. 1) that the atomic nucleus has infinite mass and, therefore, remains at rest We shall now correct for this. Calculating to begin with, in c.g.s. units, let M be the mass and ξ1 η1 ζ1 the coordinates of the nucleus, the quantities with index 2 referring to the electron of mass m. The HAMILTONIAN of the system is

    and SCHRÖDINGER’S equation is

    is the LAPLACian in the configuration space of the electron. u′ depends on the center of mass

    and the relative coordinates

    we have

    where

    is the reduced mass. (5.1) can be separated by means of a solution of the form

    The motion of the center of mass of the atom is governed by the equation

    while for the relative motion of the electron the following equation holds:

    (5.5) differs from (1.1′) only in that μ appears in the place of m. Thus, we need only to alter the atomic units defined in the Introduction in order to take into account the motion of the nucleus. Adopting μ as the new atomic unit of mass, the previously used unit of energy becomes multiplied by

    where A is the atomic weight of the nucleus. In terms of the new atomic units SCHRÖDINGER’S equation again assumes the old form (1.1). In terms of the new units the energy of the n-th discrete state of a hydrogen-like ion is again given by BALMER’S formula (2.11); in terms of the old units, which we shall retain in general, we accordingly get:

    According to (5.7) the absolute value of the energy, as a consequence of the motion of the nucleus, decreases with decreasing nuclear mass. (The unit of length a is increased by a factor of m/μ, i.e. the electrons are on the average more distant from a light nucleus than from a heavy nucleus of the same charge.)

    This effect of nuclear motion has been of importance historically both for the detection of isotopes, in particular of deuterium²⁷, and for the spectroscopic determination of the atomic mass of the electron. Consider, for instance, the various fine structure components of the Hα line in hydrogen and in deuterium (Z − 1, atomic mass about 2). After applying small corrections for hyperfine structure and relativistic effects, the wave numbers of each line for hydrogen and for deuterium differ from the wave number for infinite nuclear mass only by a multiplicative factor of form (5.6). In (5.6), M is replaced by the proton and deuteron mass, respectively, for hydrogen and deuterium. The difference in wave numbers vH and vD for the two isotopes is then given by

    where MH, MD are the mass of a neutral hydrogen and deuterium atom, respectively (including the electron mass).

    The difference in wave number (5.8), has been measured by various authors²⁸ to an accuracy of about 1 in 5000. To get a feeling for the order of magnitude of the effect, the difference in wavelength of the Hα line in hydrogen (λ = 6560 Å) and in deuterium is about 1.75 Å, or about one-third the doublet separation of the D-lines in sodium. The values of MH and MD are known very accurately from mass spectroscopy and from data on nuclear reactions. In physical atomic mass units (referred to O¹⁶) they are MH = 1.008142 a.m.u., MD = 2.014737 a.m.u. From (5.8) and the experimental values for (vD − vH) one can then calculate the atomic mass of the electron to an accuracy of about 1 in 5000. Actually the electron’s atomic mass is known more accurately from other types of experiments, in particular measurements of the cyclotron frequency of an electron and a proton in a constant magnetic field, giving directly the proton-electron mass ratio. The best non-spectroscopic value²⁹ for the electron mass is

    The spectroscopic value for m, obtained by COHEN, is larger than the more accurate one given in (5.9) by about 1 part in 2500, i.e., by slightly more than its experimental error.

    The relation between the RYDBERG constant RH for an actual hydrogen (or deuterium) atom and R∞, Eq. (2.14), is

    Of course, RH and RD are the quantities which are measured directly and R∞ is derived from them with the help of (5.9).

    6. Separation of SCHRÖDINGER’S equation in parabolic coordinates³⁰. SCHRÖDINGER’S equation for an electron moving in any central force field can always be separated in spherical polar coordinates. If the central field is of the COULOMB type, then a separation can also be carried out in parabolic coordinates. This alternative is connected with the degeneracy of the eigenvalues belonging to like principal and different orbital quantum numbers (cf. Sect. 2). A separation in parabolic coordinates turns out to be useful in the treatment of all kinds of perturbation problems in which a particular direction in space is distinguished by some external force, e.g., STARK effect, photo-electric effect, COMPTON effect, and collision of electrons.

    α) Discrete spectrum. The parabolic coordinates ξ, η, φ are defined through the relations³¹

    The surfaces ξ = const and η = const are paraboloids of revolution about the z-axis having the nucleus at the origin (x = y = z = 0) as focus. The coordinate system is orthogonal. The element of arc is given by

    and the volume element by

    From (6.2) follows the expression for the LAPLACian operator, viz.

    We deal with SCHRÖDINGER’S equation by setting

    Multiplying the differential equation by (ξ + η) and carrying out the separation we obtain

    and also an exactly equivalent equation for u2(η). A procedure analogous to that in Sect. 4 leads us to conclude that u1, behaves as for large ξ and as for small ξ. We put

    so that

    On comparison with (3.45) it becomes apparent that the solutions of this equation are

    where

    must be a non-negative integer (in the case of real ε) if f1 is to remain finite for large ξ. A corresponding result may be obtained for f2. Finally, putting

    Fig. 8. Charge distribution of the state n1 = 2, n2 = 0, m = 1 in parabolic quantization (according to F. G. SLACK): The figure shows a cross-section through the atom, the nucleus being at the center of the coordinate system. The curves are lines of constant charge density, where charge density means the charge in a circular ring with the quantization direction (z-axis) as axis. Note the strong concentration of charge towards positive values of z

    and solving (6.8) for ε, we obtain our previous energy formula (2.10), viz.,

    The degree of degeneracy of the n-th eigenvalue is, as it must be, the same as in our previous calculation in polar coordinates. If m is fixed, n1 can assume the n m values 0, 1, …, n m − 1. m itself can go from 0 to n − 1, the non-zero values having to be counted twice because one can choose either the plus of the minus sign in of (6.5). Thus, one arrives again at exactly n² different eigenfunctions.

    We must also normalize the eigenfunctions. Since the volume element is given by (6.3), we require

    The value of the integral can be taken from (3.16). The normalized eigenfunction becomes

    These eigenfunctions, contrary to the eigenfunctions in polar coordinates, are asymmetrical with respect to the plane z = 0. For n1 > n2, the larger portion of the charge distribution of the electron lies on the positive side of z for n1 < n2, on the negative side of z. This is best seen by examining the eigenfunctions for very large distances from the nucleus, i.e., for large arguments of the LAGUERRE functions. For large behaves [cf. (3.7)] as μ, and in view of the definitions (6.1) of parabolic. coordinates we have

    Fig. 8 gives contours of constant charge density for the state n = 3, n1 = 2, n2 = m = 0. The large eccentricity in the charge distribution is quite evident.

    The parabolic eigenfunctions can, naturally, be built up from the eigenfunctions in polar coordinates; e.g., for n = 2, n1, n2 = m = 0 we have

    in view of (3.7), (3.21), (1.8) and (6.1).

    Generally, any of the (n m) wave functions in parabolic coordinates for a fixed value of n and m (and fixed sign in is a linear superposition of the (n m) spherical harmonics wave functions for the same values of n and m (and sign). For the non-degenerate ground state with n = 1 (n1 = n2 = m = 0) the parabolic and spherical harmonics wave functions are identical.

    β) Continuous spectrum. For the continuous spectrum, with positive energy E, we find that

    is again a pure imaginary and

    are complex. λ can assume continuously all values from −∞ to +∞. The LAGUERRE functions with the complex indices n1 + m and n2 + m can again be represented by the integrals (4.6) and (4.7), from which a series expansion corresponding to (4.8) and an asymptotic representation corresponding to (4.10) may be derived. If the normalization is performed in the k and λ scales one obtains³²:

    where

    when m is even.

    when m is odd.

    The path of integration in f consists, exactly as in (4.7), of a simple loop around the two branch points (cf. Fig. 6). At large distances from the atom u behaves like a spherical wave, i.e., u falls off³³ as 1/r.

    γ) Eigenfunctions which behave asymptotically like plane waves. RUTHERFORD’S scattering formula³⁴. In the theory of scattering of electrons and of other charged particles by bare nuclei it is convenient to construct a wave function which behaves asymptotically like an incident plane wave, with amplitude independent of r, plus an outgoing spherical wave. Such a wave function was first discussed by GORDON.

    We try a wave function³⁵ of the form

    where We shall see that such a wave function has the required asymptotic behavior. Substituting (6.16) into the wave equation, we find that F(η) is indeed a function of η only and satisfies the equation

    Comparison of (6.17) with (3.43) shows that F(η) is, except for a normalization factor c, a LAGUERRE function

    The solution u, (6.16) with (6.18), is mathematically similar to (6.15), but for a complex value of λ,

    F(η) can also be expressed in terms of a (confluent) hypergeometric function. ADart from a normalization constant it is

    where n′ = Z/k.

    Normalizing the wave function to unit charge density at large distance from the nucleus, we obtain the integral representation for u,

    in which the path of integration, as before and exactly as in (4.7), consists of a simple loop going around the branch points In view of (6.1), u is

    represented asymptotically by

    where σn′ = arg Г(1 + in′) is the complex phase of the Г-function.

    The first term in (6.21) represents a plane wave incident in the z direction which is slightly modified by the COULOMB potential of the nucleus. Its amplitude does not depend on the separation r between electron and nucleus. The amplitude of the second term, on the other hand, is inversely proportional to r and thus represents a spherical wave; that the spherical wave is outgoing may be seen by including in (6.21) the time factor eiEt. The spherical wave is necessarily linked to the plane wave and represents the scattering of the electron by the nucleus.

    Since the amplitude of the incident wave is unity and the velocity of each electron is k (atomic units), it is clear that k electrons per unit time enter the region of interaction with the nucleus through unit area of a surface which is perpendicular to the z-axis and is located at a great distance from the nucleus. The number of electrons scattered into the solid angle per unit time, i.e., the number leaving the field of force of the nucleus through an element of area r² of a distant spherical surface per unit time, is given, in view of (6.21). by

    In the above, is the angle of deviation of the electrons caused by scattering. The coefficient for scattering at the angle thus becomes:

    S² has the dimensions of an area and is measured in atomic units (a²). To change to c.g.s. units we must put

    where E is the energy of the incident particles in c.g.s. units and thus (cf. Introduction) :

    This is the well-known scattering formula of RUTHERFORD.

    It is often useful to normalize (6.20) and (6.21) in some particular one of a number of different ways. Most frequently the normalization is such that one particle is incident on unit area per unit time; this requires multiplication of (6.20) and (6.21) by where v is the velocity (k atomic units).

    SOMMERFELD employs as eigenfunctions the system of functions

    in his theory of the continuous X-ray spectrum, k is a vector of variable direction and magnitude. The asymptotic formula for large r is

    The system uk makes the intuitive meaning of the eigenfunctions particularly clear. uk is a simple plane wave incident in the k direction plus an associated scattered wave. The eigenfunction (6.26) is normalized in the k scale. Expanding an arbitrary function of the spatial coordinates in terms of the system (6.26)³⁶ one has

    and the coefficients in the expansion are given by

    Finally, uk can be normalized also per energy interval E and element of solid angle . is the element of solid angle into which the vector k points and atomic units. Thus, This change in normalization requires a division of (6.25) and (6.26) by

    We have only discussed wave functions in the continuum which behave asymptotically like a plane wave plus spherically outgoing waves. For a physical problem involving only scattering by a pure COULOMB field, only these wave functions will occur. But in problems where the matrix element of some operator is required between an arbitrary initial state of the electron and a final state in the continuum, another set of wave functions is useful³⁷. These other wave functions behave asymptotically like a plane wave plus spherical incoming waves. They

    Enjoying the preview?
    Page 1 of 1