Quantum Orbits. Quantum Theory for the Computer Age Unit 9. Diving orbit. Caustic. for KE/PE =R=-3/8. for KE/PE =R=-3/8. p"...

Similar documents
The Hydrogen Atom. Dr. Sabry El-Taher 1. e 4. U U r

Time part of the equation can be separated by substituting independent equation

The Hydrogen Atom. Nucleus charge +Ze mass m 1 coordinates x 1, y 1, z 1. Electron charge e mass m 2 coordinates x 2, y 2, z 2

Physics 342 Lecture 23. Radial Separation. Lecture 23. Physics 342 Quantum Mechanics I

The 3 dimensional Schrödinger Equation

1 Reduced Mass Coordinates

PHYS 3313 Section 001 Lecture # 22

which implies that we can take solutions which are simultaneous eigen functions of

20 The Hydrogen Atom. Ze2 r R (20.1) H( r, R) = h2 2m 2 r h2 2M 2 R

Quantum Theory of Angular Momentum and Atomic Structure

Quantum Mechanics in Three Dimensions

(3.1) Module 1 : Atomic Structure Lecture 3 : Angular Momentum. Objectives In this Lecture you will learn the following

Chem 467 Supplement to Lecture 19 Hydrogen Atom, Atomic Orbitals

4 Power Series Solutions: Frobenius Method

The Hydrogen Atom Chapter 20

ONE AND MANY ELECTRON ATOMS Chapter 15

Outline Spherical symmetry Free particle Coulomb problem Keywords and References. Central potentials. Sourendu Gupta. TIFR, Mumbai, India

C/CS/Phys C191 Particle-in-a-box, Spin 10/02/08 Fall 2008 Lecture 11

Quantum Mechanics: The Hydrogen Atom

MITOCW watch?v=fxlzy2l1-4w

8.1 The hydrogen atom solutions

One-electron Atom. (in spherical coordinates), where Y lm. are spherical harmonics, we arrive at the following Schrödinger equation:

The Central Force Problem: Hydrogen Atom

IV. Electronic Spectroscopy, Angular Momentum, and Magnetic Resonance

quantization condition.

The Hydrogen Atom. Chapter 18. P. J. Grandinetti. Nov 6, Chem P. J. Grandinetti (Chem. 4300) The Hydrogen Atom Nov 6, / 41

2m r2 (~r )+V (~r ) (~r )=E (~r )

Lecture 4 Quantum mechanics in more than one-dimension

Angular momentum. Quantum mechanics. Orbital angular momentum

1 Commutators (10 pts)

Schrödinger equation for central potentials

Physics 228 Today: Ch 41: 1-3: 3D quantum mechanics, hydrogen atom

The Hydrogen atom. Chapter The Schrödinger Equation. 2.2 Angular momentum

Lecture 4 Quantum mechanics in more than one-dimension

df(x) = h(x) dx Chemistry 4531 Mathematical Preliminaries Spring 2009 I. A Primer on Differential Equations Order of differential equation

1.4 Solution of the Hydrogen Atom

Chemistry 532 Practice Final Exam Fall 2012 Solutions

Physics 606, Quantum Mechanics, Final Exam NAME ( ) ( ) + V ( x). ( ) and p( t) be the corresponding operators in ( ) and x( t) : ( ) / dt =...

Introduction to Quantum Mechanics (Prelude to Nuclear Shell Model) Heisenberg Uncertainty Principle In the microscopic world,

Quantum Mechanics Solutions

4/21/2010. Schrödinger Equation For Hydrogen Atom. Spherical Coordinates CHAPTER 8

Quantum Mechanics-I Prof. Dr. S. Lakshmi Bala Department of Physics Indian Institute of Technology, Madras. Lecture - 21 Square-Integrable Functions

Lecture 10. Central potential

atoms and light. Chapter Goal: To understand the structure and properties of atoms.

1 r 2 sin 2 θ. This must be the case as we can see by the following argument + L2

Orbital Angular Momentum of the Hydrogen Atom

H atom solution. 1 Introduction 2. 2 Coordinate system 2. 3 Variable separation 4

Brief review of Quantum Mechanics (QM)

Quantum Mechanics in 3-Dimensions

Atomic Systems (PART I)

Mathematical Tripos Part IB Michaelmas Term Example Sheet 1. Values of some physical constants are given on the supplementary sheet

Solved radial equation: Last time For two simple cases: infinite and finite spherical wells Spherical analogs of 1D wells We introduced auxiliary func

A few principles of classical and quantum mechanics

Quantum Physics III (8.06) Spring 2007 FINAL EXAMINATION Monday May 21, 9:00 am You have 3 hours.

Quantum Numbers. principal quantum number: n. angular momentum quantum number: l (azimuthal) magnetic quantum number: m l

3. Quantum Mechanics in 3D

We now turn to our first quantum mechanical problems that represent real, as

Schrödinger equation for central potentials

Lectures 21 and 22: Hydrogen Atom. 1 The Hydrogen Atom 1. 2 Hydrogen atom spectrum 4

Lecture 11 Spin, orbital, and total angular momentum Mechanics. 1 Very brief background. 2 General properties of angular momentum operators

Physics 6303 Lecture 11 September 24, LAST TIME: Cylindrical coordinates, spherical coordinates, and Legendre s equation

ψ s a ˆn a s b ˆn b ψ Hint: Because the state is spherically symmetric the answer can depend only on the angle between the two directions.

P3317 HW from Lecture and Recitation 7

CHEM-UA 127: Advanced General Chemistry I

d 1 µ 2 Θ = 0. (4.1) consider first the case of m = 0 where there is no azimuthal dependence on the angle φ.

1 Schroenger s Equation for the Hydrogen Atom

Representation theory and quantum mechanics tutorial Spin and the hydrogen atom

r R A 1 r R B + 1 ψ(r) = αψ A (r)+βψ B (r) (5) where we assume that ψ A and ψ B are ground states: ψ A (r) = π 1/2 e r R A ψ B (r) = π 1/2 e r R B.

Second quantization: where quantization and particles come from?

CHAPTER 6 Quantum Mechanics II

Phys 622 Problems Chapter 5

Physics 342 Lecture 26. Angular Momentum. Lecture 26. Physics 342 Quantum Mechanics I

Legendre Polynomials and Angular Momentum

Self-consistent Field

Physics 227 Exam 2. Rutherford said that if you really understand something you should be able to explain it to your grandmother.

Welcome back to PHY 3305

eff (r) which contains the influence of angular momentum. On the left is

Nuclear Shell Model. Experimental evidences for the existence of magic numbers;

CHAPTER 8 The Quantum Theory of Motion

Multielectron Atoms.

Total Angular Momentum for Hydrogen

Physics 113!!!!! Spring 2009!! Quantum Theory Seminar #9

1 Solutions in cylindrical coordinates: Bessel functions

Probability and Normalization

Deformed (Nilsson) shell model

Sparks CH301. Quantum Mechanics. Waves? Particles? What and where are the electrons!? UNIT 2 Day 3. LM 14, 15 & 16 + HW due Friday, 8:45 am

( ) = 9φ 1, ( ) = 4φ 2.

MITOCW ocw lec8

An Introduction to Hyperfine Structure and Its G-factor

Atomic Structure and Atomic Spectra

Chem 442 Review for Exam 2. Exact separation of the Hamiltonian of a hydrogenic atom into center-of-mass (3D) and relative (3D) components.

Quantum Mechanics in Three Dimensions

Angular momentum & spin

PHYS 502 Lecture 8: Legendre Functions. Dr. Vasileios Lempesis

Physics 342 Lecture 22. The Hydrogen Atom. Lecture 22. Physics 342 Quantum Mechanics I

Single Electron Atoms

Potential energy, from Coulomb's law. Potential is spherically symmetric. Therefore, solutions must have form

Addition of Angular Momenta

QUANTUM MECHANICS AND ATOMIC STRUCTURE

QUANTUM MECHANICS Intro to Basic Features

Transcription:

W.G. Harter Coulomb Obits 6-1 Quantum Theory for the Computer Age Unit 9 Caustic for KE/PE =R=-3/8 F p' p g r p"... P F' F P Diving orbit T" T T' Contact Pt. for KE/PE =R=-3/8 Quantum Orbits

W.G. Harter Coulomb Obits 6- Unit 9 Quantum Orbits Quantum theory of particle orbits has a long history beginning with the Bohr atomic Coulomb orbits introduced in Chapter 5 and D and 3D oscillator orbits introduced in Chapter and shown to have U(3) and U(3) symmetry, respectively. Both are helped a lot by the R(3) rotational states and wavefunctions introduced in Chapter 3. The 3D Coulomb orbital problem has a higher symmetry R(4) that is something like the U(3) symmetry of the 3D isotropic oscillator. This connection will be explored. W. G. Harter Department of Physics University of Arkansas Fayetteville Hardware and Software by HARTER-Soft Elegant Educational Tools Since 001

W.G. Harter Coulomb Obits 6-3 QM for AMOP Ch. 6 Coulomb Orbits W. Harter The Bohr atomic Coulomb problem has nicely closed analytic solutions for both the classical and quantum versions. This is not an accident but due to symmetry. Like the 3D isotropic oscillator the 3D Coulomb orbital problem has a higher symmetry R(4) that is something like the U(3) symmetry of the 3D isotropic oscillator. This quantum symmetry connection will be explored and used to better understand the quantum orbitals and they will be compared with the classical orbits.

W.G. Harter Coulomb Obits 6-4 Unit 9 Quantum Orbits...5 Chapter 6 Coulomb Orbits...5 6.1 Introduction to Radial Hamiltonians...5 (a) Kinetic energy operator in polar form...6 (b) Angular-radial separation...8 (c) Radial equations and solution...10 1. Potential-free case...10. Coulomb potential...10 (d) Radial wavefunctions...15

W.G. Harter Coulomb Obits 6-5 Unit 9 Quantum Orbits Chapter 6 Coulomb Orbits 6.1 Introduction to Radial Hamiltonians The ability to separate multi-dimensional equations into independent one-dimensional parts has generally been the prerequisite to finding exact analytic solutions for physics problems. This applies to classical equations of motion as well as the quantum mechanical equations of Schrodinger or Heisenberg. Further, such separation often depends on having some kind of symmetry which may or may not be obvious. Spherical rotational symmetry R(3) which has been exploited in the preceding sections is one of the more obvious ones. Closely related to R(3) is the U() symmetry of the isotropic two-dimensional oscillator. For systems with spherical symmetry, the separation of the angular parts from the radial parts is to be expected. In the case of a single orbiting particle, such as an electron around a nucleus, the wave function for eigenstates has the simple product form n,,m (r,, ) = R n (r) Y m (, ) where the angular part has already been given in closed analytic form in Chapter 3 by using the properties of rotational or R(3) symmetry operators. Now we consider examples of possible radial functions R n (r). The radial functions will depend on the form of the radial potential function V(r), and only for certain special potentials will symmetry analysis be helpful. Fortunately, two of the most common potentials, the harmonic oscillator V(r)=kr / and the Coulomb potential V(r)=k/r have high symmetries which allow symmetry analysis to aid in there analysis, as well. The three-dimensional oscillator has an obvious U(3) symmetry analogous to the U() symmetry of the two-dimensional oscillator. The Coulomb potential has a fairly non-obvious symmetry R(4) based on the eccentricity vector conservation. For the vast majority of possible radial potentials V(r), approximate solutions of their radial differential equations seem to be the only resort. Naturally, a modern approach offers numerical solution as a viable option in all cases. So far, our only treatment of radial equations has been a brief analysis of the radial equation for the isotopic two-dimensional oscillator in Sec. 1.(b). That technique will be used again here for some three-dimensional examples including that of the oscillator.

W.G. Harter Coulomb Obits 6-6 (a) Kinetic energy operator in polar form For an isotropic or spherical potential that is a function V(r) of radius alone, the separation of this degree of freedom from the angular ones is an obvious consequence of choosing polar coordinates. x = r cos sin, y = r sin sin, z = r cos (6.1.1) From this follows the Jacobian transformation matrix. x x x r y y y cos sin rcos cos rsin sin = sin sin rsin cos rcos sin (6.1.a) r z z z cos r sin 0 r The inverse (Kajobian) matrix is useful, too. It is inverse by chain rules: r x x x r y y y r cos sin sin sin cos z cos cos sin cos sin = z r r r sin cos 0 z rsin rsin x a r µ µ r µ x b = x a x b = b a, etc. (6.1.b) A transpose of the latter gives Cartesian momentum operators in terms of radial-polar angle derivatives. r cos cos sin x x x x r cos sin r rsin r r sin cos cos = = sin sin (6.1.3a) y y y y r rsin r sin z z z z cos 0 r The inverse of this is a Jacobian differential operator relation based on the transpose of (6.1.a). x y z r r r r x x y z cos sin sin sin cos x = = rcos cos rsin cos r sin (6.1.3b) y y x y z r sin sin rcos sin 0 z z The radial part of the above, multiplied by r, is worth noting. r r = rcos sin x + rsin sin y + rcos z = r r = r (6.1.3c)

W.G. Harter Coulomb Obits 6-7 Classical Lagrangian mechanics begins with the first differential vector. dx a = xa µ r µ drµ = xa a x xa dr + d + r d, where: xa = { x, y,z} Its square arc-length or metric has three terms (instead of nine) because polar coordinates are orthogonal. ( ds) = ( dx a ) = x a x a a a,µ, r µ r dr µ dr = ( dr ) + r ( d ) + r sin ( d ) This gives a velocity-squared which, multiplied by half the mass, is the non-relativistic kinetic energy. T = 1 M s = r + r + r sin Hamiltonian mechanics uses momenta p µ that use inverse (Kajobian) relations like (6.1.3). p a = M x a = rµ µ x a p µ = r x a p r + x a p + x a p, where: x a = dx dt, dy dt, dz (6.1.4) dt Again, this gives a three-term expression for classical kinetic energy using polar momenta. T = 1 p M ( a ) = r µ r a µ x a x a p µ p = pr M + p Mr + p Mr sin (6.1.5a) This is consistent with the fundamental Lagrangian definitions of momentum. p r = T r = M r, p = T = Mr, p = T = Mr sin, (6.1.5b) The quantum Hamiltonian is simplified by the square L orbital momentum in Cartesian form. L = L L = ( r p) ( r p) = abc r b p c ade r d p e a,b,c,d,e = b,c,d,e ( bd ce be cd )r b p c r d p e (6.1.6) = ( r b p c r b p c r b p c r c p b ) b,c The Levi-Civita identity a abc ade = bd ce be cd is used followed by commutation relations. r a p b = p b r a + i ab, p b r a = r a p b i ab. (6.1.7) This is rearranged to give r p = r b p b terms like (6.1.3c) wherever possible. ( ) ( ) L = r b r b p c p c ir b p c bc r b p c p b r c + ir b p c bc = r b r b p c p c r b p b p c r c b,c b,c = ( r b r b p c p c r b p b r c p c + ir b p b ) b,c (6.1.8)

W.G. Harter Coulomb Obits 6-8 The result is concise expression involving the radial coordinate and momentum only. L = ( r r)( p p) ( r p)( r p) + i r p ( ) = r p ( rp r )( rp r ) + i( rp r ) The quantum kinetic energy reduces to the following. The classical T emerges if we ignore the tiny 0. ( rp r )( rp r ) i rp r T = p M = ( ) Mr + L Mr = p r ( rp r ) ip r + L p Mr Mr r 0 M + L Mr The Jacobian relation (6.1.3c) gives a representation for rp r equal to (i/)r r. r p = rp r = i r r = i r r = i r This gives the complete quantum kinetic energy operator in polar coordinates. T = M = ( r r )( r r ) + r r M = r r + rr M r + L Mr ( ) + L ( ) r + L Mr (6.1.9) (6.1.10a) = 1 M r r r r Mr This may be compared with the polar coordinate Laplacian form arising from (6.1.3a). T = 1 M r r ( r r ) Mr sin sin ( ) Mr sin (6.1.10b) From this we infer the coordinate differential form for the square-angular momentum. L = sin ( sin ) + sin (6.1.10c) (b) Angular-radial separation The classical Hamiltonian for a radial potential V(r) separates nicely by assuming angular momentum conservation in the kinetic part (6.1.9). H = p M +V ( r) p r 0 M + L Mr +V ( r) (6.1.11) The result is a 1-D potential V(r) plus a 1/r effective potential known as the centrifugal barrier term. The quantum Hamiltonian Schrodinger energy eigenvalue equation has a purely radial form. H = M 1 +V (r) = E = M r r r r + L Mr (6.1.1) This separates nicely by using the eigenvalues (+1) from (3.1.10b)) for L acting on spherical harmonics m=y m(), for which we have analytic expressions (3.1.15) and (3.1.16).

W.G. Harter Coulomb Obits 6-9 L m = L Y m ( ) = ( ) sin sin Y m + Y m sin ( ) = +1 ( )Y m ( ) (6.1.13a) Then the equation for the radial part R n (r) of the product wave function n,,m (r,, ) = R n (r) Y m (, ) (6.1.13b) is as follows. 1 d M r dr r dr dr + ( +1) Mr + V (r) E R = 0 (6.1.13c) Before proceeding with solutions for various radial potentials of this equation, it is instructive to note that a polar coordinate separation is possible in some cases even if the potential has angular dependence as well as radial, provide the angular dependency has the following form. 1 r r r r + 1 r sin sin + 1 r sin + u vr ( ) ( ) w ( ) r r sin = 0 (6.1.14a) The energy and potential functions are in natural atomic units here. = M M E, v(r) = V (r), etc. (6.1.14b) Substituting (6.1.13b) and multiplying through by r /RY gives the following. 1 R r r R r + r 1 Y ( vr ( )) + sin Y sin u ( )Y + 1 Y sin w ( ) r sin Y = 0 The radial part must equal a first separation constant k 1 since it is independent of the angular part. An ordinary radial differential equation like (6.1.13c) results. The separation constant is k 1 =(+1). 1 d r dr r dr dr + vr ( ) k 1 r R = 0 (6.1.15a) Also the angular part is a constant, too. It is minus the separation constant. A product wave function Y(, )= P()Q() (6.15.5b) is assumed in order to separate the rest. Q P k 1 + sin POsin u ( )PQ + P Q sin w ( ) r sin PQ = 0 Divide by PQ/sin and rearrange so that a second separation constant k is warranted. sin P sin P ( u ( ) k 1 )sin + 1 Q Q w ( ) = 0 The result is two more ordinary differential equations. sin d dp sin d d (( u( ) k1 )sin + k ) P = 0 (6.15.5c) (6.1.15d)

W.G. Harter Coulomb Obits 6-10 The Q equation is a 1-D wave equation. For w=0 it reduces to Bohr orbital plane-wave exponentials. Q=Q m () = A e im + A e -im (6.1.16) The separation constant is k = m according to (6.1.15d). It may be substituted in the P=P(q) equation (6.1.15c).to give the following. sin d dp sin d d (( u( ) k1 )sin m ) P = 0 (6.1.17a) For u=0 this reduces to Legendre's equation. We use separation constant is k 1 =(+1). sin d dp sin d d + ( ( +1 )sin m ) P = 0 5.1.17b) Compare this to angular eigenfunction equation (6.1.13c) which has not yet separated and angles. (c) Radial equations and solution We now consider some special cases of the radial differential equation (6.1.13c) when wellknown potentials are used. The simplest cases involve no potential at all. 1. Potential-free case With V(r)=0 the radial equation (6.1.13c) becomes the following in atomic energy units. 1 d r dr r dr +1 dr ( ) r R = 0 (6.1.18) If the energy also happens to be zero (E=0) then this is the radial part of the Laplace equation =0. d dr r dr dr ( +1)R = 0 (6.1.19a) The solutions for the latter are the any combination of the following powers. R = R ( r) = Ar + B 1 r +1 (6.1.19b) This is consistent with the multipole expansions (3.4.10), all of which are solutions to the Laplace equation for electrostatic potentials in a vacuum around charges. However, the radial functions for quantum particles in free space are quite different. In general the solutions of (6.1.18) can be given in terms of spherical Bessel functions. R = R ( r) = Aj ( r) + Bh ( r) These important functions will be discussed in connection with scattering theory.. Coulomb potential With v(r)=k/r the radial equation (6.1.13c) becomes the following in atomic energy units. 1 d r dr r dr +1 dr ( ) r + k r R = 0 (6.1.0a)

W.G. Harter Coulomb Obits 6-11 We shall here consider just the bound state case which requires a negative value for both the Coulomb constant (k=- k makes the potential attractive) and the energy (=- makes the wave bounded in space.) 1 d r dr r dr +1 dr ( ) r k r + R = 0 (6.1.0b) For the Coulomb potential, several re-scaling factors are conventional and convenient. First, a radial scale R(r) = X (r) r, dr dr = 1 dx r dr - X(r) r, d R dr = 1 d X r dr dx X (r) r - dr r 3 makes the radial integral r R dr become simply X dr and gives the standard form derived below. From d R dr + dr ( +1 )R r dr r + k R R = 0 r we get an X-wave equation which has no first-derivative term. 1 d X r dr ( +1) X r r + k X r r X r = 0 or: d X +1 dr ( ) r X + k X X = 0 (6.1.0c) r Finally, most Coulomb treatments use atomic units of distance and energy for radius r=a and E =E R. d X +1 d ( ) X + X X = 0 (6.1.0d) where /a = M k / and /a = M E / are given in terms of Rydberg energy E R and Bohr radius a. r = a = M k = 4 0 Me Z (6.1.0e) E = E R = Ma (6.1.0f) One should pause and realize that this equation has been solved on the average of at least once a day for over 75 years; that is, roughly 7,375 times. (And, here goes 7,376!) So, by now everyone thinks they have a pretty good scheme of units. The Bohr radius a= 0.58 Å was discussed in (5.6.3). The Rydberg energy for atomic number Z=1 (Hydrogen) is E R = 13.6eV. To begin solving such an equation it helps to check out how it behaves for large and small radii. ( for ): d X d 0 + 0 X = 0 implies: X ~ e (6.1.1a) ( for 0): d X +1 d ( ) X +* 0 = 0 implies:x ~ +1 (6.1.1b)

W.G. Harter Coulomb Obits 6-1 This helps decide what factors to take out of the trial solution in order to possibly get a residual polynomial solution. Recall that the oscillator has an approximate e -ax solution at large distances, and that turns out to ride on every exact solution including the ground state (for which it is the whole story!). The same happens here as we will see; an e - factor will grace every bound-state Coulomb wavefunction. So the trial solution taken here will be the following expansion around =0. X = e µ f = e f +µ (6.1.) =0 =0 With the factor µ we are pretending we don't know that (6.1.1b) suggests a factor +1, and derive this later. Now we have the unenviable job of organizing the derivatives of the trial X. dx d = e +µ + e ( + µ )f f =0 =0 +µ1 d X d = e f +µ e ( + µ )f +µ1 +e ( + µ 1) ( + µ ) f +µ =0 =0 =0 To save space let us label (+µ+a) by a and write out enough terms to bring the sum power up to 0. d X d = e f 0 0 f 1 + 1 0 f where: a = + µ + a =0 =0 =0 d X d = e f 0 =0 +e µ f 0 µ1 1 f +1 0 +e + µ µ1 ( ) f 0 µ + µ +1 =0 ( )µ f 1 µ1 + 1 f + 0 =0 In this way all the terms of the radial equation (6.1.0d) are assembled power-by-power. Let L=(+1).

W.G. Harter Coulomb Obits 6-13 X = e f +µ = e [ f ] =0 =0 +µ X = e f +µ1 = e f 0 µ1 +µ + [ f +1 ] =0 =0 LX = Le f +µ = e -Lf 0 µ - Lf 1 µ1 + [ Lf + ] +µ =0 =0 d X d = e µ ( µ1) f 0 µ + ( µ +1)µ f 1 µ1 + +µ [ f - 1 f +1 + 1 f + ] =0 The sum of coefficients for each power must sum to zero. For the lowest power f 0 µ the sum gives µ(µ-1) = L = (+1) implies: µ = +1 or: µ =- (6.1.3) This is known as an index or indicial equation. Here we only accept the first solution (µ=+1) which also came from (6.1.1b) since the other possibility (µ =-) would cause µ to blow up at =0. A formula for coefficients f follows from zeroing the bracketed coefficients of +µ. ( 1 )f +1 = ( 1 L)f + where: L=(+1), or: 1 ( ( + µ +1) )f +1 = (( + µ +1) ( + µ + ) ( +1 ) f + (6.1.4) This is called a recursion relation. Let us replace by 1 and use indicial solution (µ=+1). ( ( + µ )1) f +1 = ( + µ )( + µ +1) ( +1) f = ( ( + +1)1 ) ( + +1) ( + + ) ( +1) f ( + +1 = ( )1) (6.1.5a) ( + +1) ( + + ) f Demanding that this recursion terminate for some integer = n, that is, the radial wave is a polynomial of order is n, leads to the Coulomb quantization condition on the energy parameter. 1 f n +1 = 0 implies: ( n + +1)1 = 0 or: = (6.1.5b) ( n + +1) From the energy definition (6.1.0f) follows the famous Rydberg energy level formula. 1 E = ( n + +1) E R = 1 ( n + +1) Ma = 1 N E R (6.1.5c) The energy is a function of only a single principle quantum number N. N = n + + 1 (6.1.5d) N is a sum of radial quantum number n and the total angular quantum number. Since neither n nor may be negative the two quantum numbers are limited by each other in a given energy level belonging to a particular principle N, that is n is between zero and N--1 and is between zero to N-n- 1. The n-level structure for n(=0)=ns, n(=1)=np, n(=)=nd, n(=3)=nf, and n(=4)=ng levels are sketched in Fig. 6.1.1 below. Recall that each orbital level has a +1 degeneracy due to magnetic

W.G. Harter Coulomb Obits 6-14 or z-momentum m between and -. Together, this makes each (N = n + + 1)-level have a degeneracy of N. Fig. 6.1.1 Array of Coulomb states with their quantum numbers and degeneracy. The recursion relation (6.1.5a) may be expressed in terms of the principle N-number using the quantization condition (6.1.5b). + +1 N + +1 1 f +1 = ( + +1) ( + + ) f = ( + +1 N 1) N ( + +1) ( + + ) f (6.1.5e) ( N ) = N ( + +1) ( + + ) f Radial wave functions have the polynomial following coefficients. f 1 = (0 n) N ( + ) f 0, f = (1 n) N ( + 3) f 1, f 3 = ( n) N 3 ( + 4) f, = (0 n)(1 n) N ( + ) ( + 3) f 0, = 3 (0 n)(1 n)( n) N 3 ( + ) ( + 3) ( + 4 ) f 0, The resulting wavefunction is an N-degree polynomial with a factor f 0 µ e - =f 0 e -/N R N ( r ) = X N a = f 0 e /N a 1 + (0 n) + ( ) where: r = a (0 n)(1 n) + N! + ( )( + 3) N (0 n)(1 n)( n) + 3! + ( )( + 3) ( + 4 ) N (6.1.6) 3

W.G. Harter Coulomb Obits 6-15 (d) Radial wavefunctions Some examples of these waves are plotted in Fig. 6.1.. The number and type of nodal surfaces for the complete Coulomb wave function R n (r) Y m (, ) are listed in the following table. n=n--1 concentric r-spheres -m - cones ("magic" angles) m -planes (azimuthal nodes) N-1 nodal surfaces of some kind (6.1.7) The principle or total quantum number N determines the number of nodal surfaces while the radial quantum number n=n--1 determines the number of radial nodes. The total angular quantum number determines the total number of angular modes and m determines the number of azimuthal nodes. Fig. 6.1. Lowest Coulomb atomic radial wavefunctions

W.G. Harter Coulomb Obits 6-16