Anda di halaman 1dari 85

Chapter 4

(Fairly) easy solution of the Schrodinger equation


for a three-dimensional rotationally invariant system
Version 8.35: August 17, 2010

Contents
4.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.2. The radial equation and mathematical properties of its solutions . . . . .
4.2.1. Bound and continuum states . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.2.2. Radial equations and the radial quantum number . . . . . . . . . . . . . . . . .
4.2.3. Energy ordering of bound states and binding energies . . . . . . . . . . . . . .
4.2.4. Energies and the Virial Theorem . . . . . . . . . . . . . . . . . . . . . . . . . .
4.2.5. Boundary conditions on the radial functions . . . . . . . . . . . . . . . . . . . .
4.2.6. The radial quantum number and nodes in the radial function . . . . . . . . . .
4.2.7. Orthogonality and nodes in the radial functions . . . . . . . . . . . . . . . . . .
4.2.8. The essential degeneracy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.3. The radial probability density . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.3.1. Bookkeeping, spectroscopic notation, and the principal quantum number . . .
4.3.2. The position probability density according to Born . . . . . . . . . . . . . . . .
4.3.3. Derivation and interpretation of the radial probability density . . . . . . . . . .
4.3.4. Integrated probability densities: (finite) shell and sphere probabilities . . . . .
4.3.5. Radial expectation values and uncertainties . . . . . . . . . . . . . . . . . . . .
4.4. The case of the disappearing derivative: The reduced radial equation . . .
4.4.1. Introducing the reduced radial equation . . . . . . . . . . . . . . . . . . . . . .
4.4.2. Boundary conditions on the reduced radial function . . . . . . . . . . . . . . .
4.4.3. Subduing the reduced radial equation through pattern matching . . . . . . . .
4.4.3.1. The structure of the reduced radial equation . . . . . . . . . . . . . .
4.4.3.2. The effective potential energy . . . . . . . . . . . . . . . . . . . . . . .
4.4.4. Orthonormality of the reduced radial function . . . . . . . . . . . . . . . . . . .
4.4.5. Summary: Why bother with the reduced radial function? . . . . . . . . . . . .
4.5. The roles of the physical and effective potential energies . . . . . . . . . .
4.5.1. The competition between the barrier term and the physical potential energy . .
4.5.2. The effective potential energy and the number of bound states . . . . . . . . .
4.6. Generic properties of the reduced radial function . . . . . . . . . . . . . . .
4.6.1. Limiting behavior of physically reasonable potential energies . . . . . . . . . .
4.6.2. Cleaning up the reduced radial equation . . . . . . . . . . . . . . . . . . . . . .
4.6.3. Generic behavior of reduced radial functions in the asymptotic limit . . . . . .
4.6.3.1. Derivation of the asymptotic reduced-radial equation . . . . . . . . .
4.6.3.2. Solution of the asymptotic reduced-radial equation . . . . . . . . . . .
4.6.4. Generic behavior of reduced radial functions in the near-zero limit . . . . . . .

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

265
267
267
268
269
269
270
271
271
273
274
274
276
276
279
280
281
281
282
283
283
284
284
285
286
286
287
288
289
290
291
291
291
293

264

4. The central-force problem in three dimensions


4.6.5. Generic behavior between the near-zero and asymptotic limits . . . . . . . . .
4.7. Qualitative solution of the radial equation . . . . . . . . . . . . . . . . . . .
4.7.1. Classically allowed and classically forbidden regions . . . . . . . . . . . . . . .
4.7.2. Classical turning points . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.8. Final thoughts: tips for potential modelers . . . . . . . . . . . . . . . . . . .
4.9. Users guide to Chap. 4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.10. Selected readings & references for Chap. 4 . . . . . . . . . . . . . . . . . . .
4.10.1. Numerical solution of the Schr
odinger equation . . . . . . . . . . . . . . . . . .
4.10.2. Programs for solving the radial equation . . . . . . . . . . . . . . . . . . . . . .
4.11. Exercises & problems for Chap. 4 . . . . . . . . . . . . . . . . . . . . . . . .
Complements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.A. Additional examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.B. Bound- and continuum-state radial functions . . . . . . . . . . . . . . . . . .
4.B.1. Bound versus continuum states . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.B.2. Boundary conditions on bound- and continuum-state radial functions . . . . .
4.C. The number and energy-ordering of bound states . . . . . . . . . . . . . . .
4.C.1. The number of bound states in a 3D attractive potential energy . . . . . . . .
4.C.2. The order of bound-state energies of a rotationally invariant system . . . . . .
4.D. The ground state of the lithium atom: a tale of three models . . . . . . . .
4.D.1. The principle of modeling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.D.2. Models of the ground state of lithium . . . . . . . . . . . . . . . . . . . . . . .
4.D.3. Models for the valence electron in lithium . . . . . . . . . . . . . . . . . . . . .
4.D.4. A radial function for the valence electron in Li . . . . . . . . . . . . . . . . . .
4.E. How to cope with an arbitrary wave function . . . . . . . . . . . . . . . . .
4.E.1. The method of eigenfunction expansion . . . . . . . . . . . . . . . . . . . . . .
4.E.2. Calculation of expectation values and related quantities . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

295
297
298
300
302
303
309
309
310
310
329
330
333
333
334
335
335
336
337
338
338
339
342
345
345
346

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

267
273
277
278
279
280
283
287
290
291
293
293
298
300
311
312
313
316
318
318
319
319
320
320
321
324
325
337
340
342

Figures
4.1. Bound and continuum state energies for a spherically symmetric potential energy. . .
4.2. Radial functions and the orthogonality integrand for s states of lithium. . . . . . . .
4.1. A spherical shell in 3D. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.2. Radial probability densities for 1s and 2s electrons in atomic lithium. . . . . . . . .
4.3. Radial probability densities for 1s electrons in several atoms. . . . . . . . . . . . . .
4.4. Sphere probability for electrons in the ground state of atomic lithium.. . . . . . . . .
4.1. Radial and reduced radial functions for s states of atomic lithium. . . . . . . . . . .
4.1. Effects of the centrifugal barrier on the effective potential energy. . . . . . . . . . . .
4.1. Large r behavior of some typical physical potentials energies. . . . . . . . . . . . . .
4.2. Asymptotic behavior of the ` = 1 reduced radial equation for boron. . . . . . . . . .
4.3. Limiting behavior of bound-state reduced radial functions for scandium. . . . . . . .
4.4. Near-origin behavior of the reduced radial equation for boron. . . . . . . . . . . . . .
4.1. Comparison of a model Li potential energy to pure-Coulomb potential energies. . . .
4.2. Classically allowed and classically forbidden regions for a reduced radial function. . .
4.1.1.The Yukawa model of the potential energy of the deuteron. . . . . . . . . . . . . . .
4.2.1.Reduced radial functions for two bound states of a Yukawa potential energy. . . . . .
4.2.2.The reduced radial functions for a bound state of the Yukawa potential energy. . . .
4.1.
The pure-Coulomb and Yukawa Potentials. . . . . . . . . . . . . . . . . . . . . . .
4.2.1.An effective potential energy and its reduced radial function. . . . . . . . . . . . . .
4.3.1.The reduced radial functions for Problem 4.3. . . . . . . . . . . . . . . . . . . . . . .
4.4.1.The reduced radial functions for Problem 4.4. . . . . . . . . . . . . . . . . . . . . . .
4.5.1.Three radial probability densities for Problem 4.5. . . . . . . . . . . . . . . . . . . .
4.6.1.Figure for Problem 4.6. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.7.1.The effective potential energy and two reduced radial functions for Problem 4.7. . .
4.9.1.Two effective potential energies for Problem 4.9. . . . . . . . . . . . . . . . . . . . .
4.12.1. A crude model of the nuclei of a diatomic molecule. . . . . . . . . . . . . . . . . . .
4.12.2. Absorption spectra for HCl. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.D.1.Spatial coordinates for the three electrons of atomic lithium. . . . . . . . . . . . . .
4.D.2. A model potential and screening function for the valence electron in an alkali atom.
4.D.3. The radial function and probability density for the 2s electron in Li. . . . . . . . .

JQPMaster

Version: 8.35

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

Printed: August 17, 2010

265
4.D.4. The radial probability density and shell probability for the 2s electron in Li. . . . . . . 343
4.D.5. Integrands of radial integrals for the 2s electron in Li. . . . . . . . . . . . . . . . . . . . 344

Tables
4.1. Radial functions labeled by the radial and principal quantum numbers. . . . . . . . .
4.2. Spectroscopic notation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.1. Notation for solution of the central-force problem in 3D. . . . . . . . . . . . . . . . .
4.2. Symmetry and separation of variables for single-particle potentials. . . . . . . . . . .
4.3. Symmetries and angular momentum commutator relations. . . . . . . . . . . . . . .
4.4. The one- and three-dimensional time-independent Schr
odinger equation. . . . . . . .
4.5. Behavior of the reduced radial function in classically allowed and forbidden regions.
4.6. The development of the stationary angular momentum eigenstates. . . . . . . . . . .
4.7. Properties of the reduced radial function for any spherically symmetric potential. . .
4.8. The structure of the reduced radial function. . . . . . . . . . . . . . . . . . . . . . .
4.9. Key properties of the reduced radial functions for bound and continuum SAMEs. .
4.10. Qualitative behavior of the radial function. . . . . . . . . . . . . . . . . . . . . . .
4.11. Properties in terms of the radial and reduced radial functions. . . . . . . . . . . .
4.2.1. The mean radius versus the radius of maximum radial probability density. . . . .
4.16.1. Data for the Yukawa potential energy . . . . . . . . . . . . . . . . . . . . . . . . .
4.D.1. Radial mean values for the 2s state of atomic lithium. . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

275
275
304
304
304
305
306
306
307
307
307
308
308
312
328
344

4.1. Introduction
You know what happens next. The parable is perfectly transparent.
But I have to tell you. I have to believe that my meaning resides not in the gross motion of
the tale, but in the tics of syntax and cadence.
A Moment at the Rivers Heart (Chiliad: A Meditation, Part Two), by Clive Barker

We now come to grips with the quantum-mechanical central-force problem in all its glory. In classical physics,
Definition (central force field) A particle is in a central force field provided the net force that acts
on the particle is directed radially (towards or away from the origin) with a magnitude that depends only
on the radial distance of the particle from the origin. Such a force can be written F = F (r)er , where er
is the radial unit vector of spherical coordinates (see Complement 1.D of Chap. 1). The magnitude F (r) is
positive for an attractive force (one directed towards the origin) and negative for a repulsive force (one
directed away from the origin).
In quantum mechanics, the concept of a force is a bit awkward, because we lack the concept of a
trajectory. So its easier to discuss the central-force problem in terms of the corresponding potential
energy. In classical physics, we can write any conservative force as F = V and thereby define a potential
energy V . For a central force F = F (r)er , the function V is independent of the polar and azimuthal angles
and : its spherically symmetric
V (r) = V (r, , ) = V (r),

the potential energy corresponding to a central force.

(4.1.1)

The same symmetry property applies in quantum mechanics, where a central potential energy defines a
rotationally invariant system, the type we studied in Chap. 3.
In that chapter, we sought solutions to the time-independent Schrodinger equation (TISE)
#
"
1 b2
rad
b
L (, ) + V (r) E E,`,m` (r, , ) = 0,
(4.1.2)
T (r) +
2
|2mr {z
}
angular KE operator

b rad is the radial kinetic-energy operator, and where the angular kinetic-energy operator
where T
2
b
L (, )/2mr2 , contains the square of the orbital angular momentum operator, b
L 2 . Symmetry led us to
JQPMaster

Version: 8.35

Printed: August 17, 2010

266

4.1. Introduction

seek separable solutions to Eq. (4.1.2):


E,`,m` (r, , ) = RE,` (r)Y`,m` (, )

(4.1.3)

We spent much of Chap. 3 learning about the angular factors in these solutions, the spherical harmonics Y`,m` (, ), and ended by deriving the radial equation

2
b rad (r) + ~ `(` + 1) + V (r) RE,` (r) = E RE,` (r),
T
2mr2

` = 0, 1, 2, . . .

(4.1.4)

Solution of this equation gives the radial factors RE,` (r) in the SAME functions (4.1.3) along with the
stationary-state energies E. These energies may correspond to bound states or to continuum (scattering)
states.
The many results of Chap. 3 that deal with orbital angular momentumcommutation relations, eigenvalue equations and their solutions, expectation values and uncertainties, and so onpertain to any singleparticle system. In this chapter well use those results to learn as much as we can about solutions of the
radial equation (4.1.4) without actually solving it. The heart of this chapter are generic properties of
radial functions for spherically symmetric potential energies (4.6). Generic properties are valuable, because, like the spherical harmonics, they pertain to all rotationally invariant systems. These generic
properties lead to procedures for solving the 3D TISE qualitatively (4.7). In addition to all this stuff,
a new theme (and problem-solving habit) appears in this chapter: modeling complicated systems. Like
symmetry, this theme will pervade the rest of this book. Our first direct mathematical assault on a specific
radial equation comes in Chap. 5, where we engage one-electron atoms.
In this chapter, well follow a path we first trod in 2D. We begin in 4.2 with a close look at the
structure of the radial equation (4.1.4) and the mathematical properties of its solutions: boundary conditions,
orthonormality, and nodes, properties we first encountered in Chap. 2. These properties establish the
foundation for understanding the physics inherent in radial functions via the radial probability density
of 4.3.
Once you qualitatively understand 3D radial functions, youll be ready to tackle the radial equation
mathematically. In 4.4 well discover that introducing a related function, the reduced radial function,
dramatically simplifies that mathematical chore. We further introduce the effective potential energy
(4.5), which yields insights into the solutions of the radial equation without having to solve that equation.
From these insights, we then develop generic mathematical and qualitative properties of all radial functions
for all rotationally invariant single-particle systems (4.6).
By approaching the radial equation and its solution this way, rather than immediately diving into its
detailed solution, youll learn how to brainstorm (Begin by brainstorming) central-force problems in advance
of solving them and how to pierce the inevitable veil of mathematics to see the underlying physics. Thats
why most of this chapter emphasizes the mathematical structure of the radial equation and how that
structure influences radial functions and energies. Such insights are precious, because, with very few
exceptions, radial equations cannot be solved without approximations.

JQPMaster

Version: 8.35

Printed: August 17, 2010

267

4.2. The radial equation and mathematical properties of its solutions


There was mystery hereyes; but it might not be beyond human understanding. He had
learned a lesson, though it was not one he could readily impart to others. At all costs, he
must not let Rama overwhelm him. That way lay failure, perhaps even madness.
Rendezvous with Rama, y Arthur C. Clarke

A lot of information about solutions of the 3D radial equation follows immediately by analogy to the solutions
of the 2D radial equation (Chap. 2). In this section well begin this generalization, remaining ever vigilant
for differences due to the additional dimension.
4.2.1 Bound and continuum states
A typical 3D potential energy supports bound states and continuum states.1 In terms of the systems
energies, the difference between bound and continuum states is profound. The energies of bound states are
quantized : they constitute a list of discrete valuesthe only values allowed by nature. In contrast, the
energies of continuum states can have any value above some minimum (see Complement 4.B.)
All potential energies well investigate will conform to a few limitations. These are not severe limitations:
they characterize almost all potential energies used in quantum mechanics:
(1) Well consider only potential energies V (r) that decay to zero with increasing distance r from the force
center at the origin. If we choose the zero of energy as in Fig. 4.1, then we require
V (r) 0
r

large-r behavior of a
typical potential energy.

(4.2.1a)

With this choice of the zero of energy, all discrete bound-state energies are negative, while all continuumstate energies are positive.2
0.5

Figure 4.1. A generic spherically


symmetric potential energy that
supports two bound states. The energies of these states are indicated by
the horizontal lines. Both energies are
negative because I chose the zero of energy so the r limit of V (r) is zero.
The continuum of allowed energies is
the shaded region, which includes all
values E > 0.

continuum
0.0

bound states

-0.5

-1.0
0

(2) Well consider only potential energies that are sufficiently attractive to support at least one bound
stationary state(Complement 4.C). With the zero of energy in Fig. 4.1, this means that the potential
energy must be negative, V (r) < 0, at least in some finite region of r.
(3) Well consider only potential energies that near the origin behaves as
1 Jargon: In one of the more quaint bits of jargon in quantum mechanics, a potential energy is said to support N bound
states if there exist N solution to its TISE with discrete energies.
2 Commentary: Physicists sometimes use model potential energies that do not decay to zero asymptotically. A prime example
is a 3D isotropic simple harmonic oscillator of frequency : this function, V (r) = m 2 r2 /2, becomes infinite in the r
limit. Such behavior is non-physical, so a model with this behavior is only useful for the study of states with energies deep in
the wellstates whose wave functions decay rapidly to zero with increasing r.

JQPMaster

Version: 8.35

Printed: August 17, 2010

268

4.2.2 Radial equations and the radial quantum number


lim r2 V (r) = 0

r0

small-r behavior of a
typical potential energy.

(4.2.1b)

The modest limitations in Eqs. (4.2.1) accommodate such physically important cases as potential energies
for atoms, molecules, and solidsincluding the pure-Coulomb potential energy of an electron in a
hydrogen atom (Chap. 5)but preclude certain mathematical pathologies that complicate solution of the
radial equation.3

4.2.2 Radial equations and the radial quantum number


For a potential energy that obeys Eqs. (4.2.1), a separate radial equation [Eq. (4.1.4), p. 266] exists for each
orbital angular momentum quantum number : ` = 0, 1, . . . .
Solution of each of these equations may yield no, one, or several bound-state functions RE,` (r). So we
need an index to label bound-state radial functions for a particular `. A sensiblethough not uniqueindex is
a radial quantum number nr . As in 2D (2.9.1), here nr is a non-negative integer : nr = 0, 1, 2, . . . , nmax
.
r
For any `, the upper limit nmax
may
be
zero,
a
finite
number,
or
even
infinity.
I
will
label
each
radial
r
function and bound-state energy as Rnr ,` (r) and Enr ,` . We learned in Chap. 3 that for each nr and `, there
are (2` + 1) SAMEs nr ,`,m` (r, , ) of the form (4.1.3), because there are (2` + 1) values of the projection
quantum number in ` m` `.
Here are the radial equations for the first three values of `; each is clearly a mathematically different
equation:
h
i
b rad + V (r) En ,0 Rn ,0 (r) = 0,
T
(` = 0),
(4.2.2a)
r
r

2
b rad + ~ + V (r) En ,1 Rn ,1 (r) = 0,
T
(` = 1),
(4.2.2b)
r
r
mr2

3~2
rad
b
+ V (r) Enr ,2 Rnr ,2 (r) = 0,
(` = 2).
(4.2.2c)
T +
mr2
The difference arises from the term ~2 `(`+1)/2mr2 , which appeared when the angular kinetic-energy operator
acted on spherical harmonics in the derivation of the radial equation (3.12).
Rule: The total number of bound-state solutions of each radial equation depends on the value of ` in that
+ 1. The total number
equation. The number of bound-states with a given ` that are supported by V (r) is nmax
r
of bound states supported by V (r) is the sum of the numbers of bound states for each `.4
Try This!

4.1. Dependence of SAME energies on quantum numbersr.


Use symmetry arguments to explain why 3D energies and radial functions do not depend on the projection quantum number m` .
3 Details: Physically reasonable spherically symmetric potential energies have the following properties: (1) In the near-zero
limit r 0 the magnitude of the potential energy, |V (r)|, either approaches a constant or approaches infinity less rapidly
than 1/r: that is, |V (r)| rp as r , where p 1. (2) In the asymptotic limit r , the magnitude |V (r)| goes to
zero at least as fast than 1/r. (3) The potential energy is continuous for all 0 < r < except, perhaps, at a finite number
of points.
4 Read on: Physicists have developed ways to predict the number of bound states supported by certain types of potential
energies; see Complement 4.C.

JQPMaster

Version: 8.35

Printed: August 17, 2010

4.2.3 Energy ordering of bound states and binding energies

269

4.2.3 Energy ordering of bound states and binding energies


Once we have solved the radial equation for a particular `, we can arrange the resulting bound-state energies so5
E0,` E1,` E2,` Enmax
,` 0,
r

(for the same `).

ordering of bound-state
energies for a given `

(4.2.3a)

Because each bound-state energy is negative, physicists often prefer to talk about the binding energy ,
which is just the magnitude
binding energy of a bound state
with quantum numbers nr and `.

nr ,` |Enr ,` | = Enr ,` > 0

(4.2.3b)

The ground state, which is the most tightly bound state, has, sensibly, the largest binding energy. All
excited states, which are less tightly bound than the ground state, have smaller binding energies. So the
order of the binding energies is
0 0,` 1,` nmax
,`
r

ordering of binding
energies for a given `

(4.2.3c)

I Warning: This ordering pertains only to bound states with the same orbital angular momentum quantum
number `.

4.2.4 Energies and the Virial Theorem


The classical Virial Theorem. In classical physics, the Virial Theorem is a statistical statement about
time-averaged values of the kinetic and potential energies of a system.6 If the potential energy has the
power-law form
V (r) rn+1 ,
power-law potential energy,
(4.2.4a)
then according to classical Virial Theorem,
T =

1
(n + 1) V ,
2

classical Virial Theorem for


a power-law potential energy,

(4.2.4b)

where an overbar signifies a classical time average (not a quantum-mechanical expectation value). For a
potential energy V (r) 1/r, [n = 0 in (4.2.4a)], Eq. (4.2.4b) reads
1
T = V,
2

classical Virial Theorem for


a 1/r potential energy.

(4.2.4c)

The 1/r case includes the all-important pure-Coulomb potential energy ,


V (r) = Z

e20
r

pure-Coulomb potential energy,

(4.2.5)

5 Details: The use of less-than-or-equal-to () rather than less than (<) allows for the possibility that an energy may be
degenerate.
6 Read on: You can find a good introduction to the classical Virial Theorem in Classical Mechanics by John R. Taylor (New
York: University Science Books 2005).

JQPMaster

Version: 8.35

Printed: August 17, 2010

270

4.2.5 Boundary conditions on the radial functions

where Z is the atomic number, and e0 is defined by e02 e2 /40 .


The quantum Virial Theorem. The quantum counterpart of this classical Virial Theorem is expressed
in terms of expectation values rather than time averages:7
Theorem (The Virial Theorem.) If the potential energy of a system is spherically symmetric,
V = V (r), then the expectation value of its kinetic energy is

1
dV
hT i =
r
(in any quantum state.)
(4.2.6a)
2
dr
If V is a 1/r potential energy, then
1
hT i = hV i
2

quantum Virial Theorem for


a 1/r potential energy.

(4.2.6b)

Equation (15.A.6b) leads to a useful relationship between the expectation value of the total energy E and
the expectation values of the kinetic and potential contributions to it:
hT i = hEi,

and

hV i = 2hEi

(for a 1/r potential energy).

(4.2.7)

Eqs. (4.2.6) and (4.2.7) pertain to any statestationary or nonstationaryof any system whose potential
energy satisfies the stated criteria.8
Try This!

4.2. The Virial Theorem for a pure-Coulomb potential


Derive Eq. (15.A.6b) from Eq. (4.2.6a). Use hydrogen-atom wave functions from Tbl. 5.7, p. 401
in Chap. 5 to verify this theorem for a few simple cases.

Try This!

4.3. The mean kinetic energy.


Explain why the quantum-mechanical Virial Theorem (4.2.6a) implies Eq. (4.2.7) for any state of a
pure-Coulomb potential.

4.2.5 Boundary conditions on the radial functions


In 3D, as in 2D (2.9.2), the number of nodes in the radial function Rnr ,` (r) is related to the boundstate energy Enr ,` which, in turn, is related to boundary conditions on Rnr ,` (r). The 3D radial equation
[Eq. (4.1.4), p. 266] is a second-order differential equation. So to specify a particular solution we need two
boundary conditions. One condition is the value of Rnr ,` (r) at the origin r = 0. The other must ensure
that Rnr ,` (r) can be normalized, which is possible only if it goes to zero in the asymptotic limit, r :9
(1) A radial function must be bounded at the origin: that is, at r = 0, it must be finite:
|Rnr ,` (0)| <

bound-state boundary
condition at the origin.

(4.2.8a)

7 Read on: The quantum Virial Theorem is more general and powerful that this brief discussion suggests. Equation (4.2.6)
is a specialization of a more general Theorem. For a lucid introduction to this theorem, see 5 of Complement GXI in CohenTannoudji et al. (1977). For the mathematical details, see 3.10 of Galindo and Pascual (1991) or 2.7.4 of Thaller (2005).

8 Details: For a 1D system with V (x) = xn for some real constant , the commutation relation
b b
H,
xb
px = 0 yields
2hT i = n hV i, which holds for any stationary state. Its 3D extension is 2hT i = hr V (r)i. For a system of N particles, the
right-hand side of the latter equation becomes the sum of expectation values hri i V (r)i over i from 1 to N , where the
subscript on the gradient operator i implies differentiation with respect to ri . For details, see 3.6 of Merzbacher (1998)
and 9.6 of Erkoc (2007). See also Problem 4.17.
9 Commentary: Notice that Eq. (4.2.8a) is a point boundary condition: a requirement on solutions at a point, r = 0. By
contrast, Eq. (4.2.8b) is a limit boundary condition: a requirment the solution must satisfy in a limit, r .

JQPMaster

Version: 8.35

Printed: August 17, 2010

4.2.6 The radial quantum number and nodes in the radial function

271

(2) A bound-state radial function must go to zero as r :


Rnr ,` (r) 0
r

bound-state boundary condition


in the asymptotic limit.

If these conditions are met, then the radial normalization integral is finite:
Z
Rn r ,` (r) Rnr ,` (r) r2 dr < .

(4.2.8b)

(4.2.9)

Only if Eq. (4.2.9) holds can we normalize an un-normalized radial function by dividing it by the square
root of the finite normalization integral (4.2.7).

Aside. Real radial function are okay.


In general, we can omit complex conjugation of the first radial function in integrals like Eq. (4.2.9) which
otherwise would require complex conjugation. We can do so because the radial equation Eq. (4.1.4) gives us
the freedom to use real functions for bound states. Each term in this equation is real, as are the boundary
conditions (4.2.8). Moreover, the radial equation is homogeneous, so its solution is indeterminate to within
a multiplicative phase factor ei , where is any real constant. This indeterminacy, which affects no
measurable quantities, lets us use only real radial functions, as Ill do throughout this book.

4.2.6 The radial quantum number and nodes in the radial function
The radial quantum number nr = 0, 1, 2, . . . not only labels solutions of the radial equation for a particular `, it also equals the number of radial nodes in Rnr ,` (r).10 Suppose V (r) supports more than
one bound state of a given ` > 0. According to our ordering scheme (4.2.3c), the ground state has the
minimum allowed energy. So the ground-state radial function has the largest curvature that is consistent
with the boundary conditions (4.2.8) (2.9.2). This function, R0,` (r), must therefore have no nodes; sure
enough, its radial quantum number is nr = 0. The radial function of next highest lying bound statethe
first excited state R1,` (r)has the next largest curvature consistent with the boundary conditions. Its
radial function therefore has one node, and its labeled by nr = 1. And so forth.
Key point! For a radial function Rnr ,` (r), the radial quantum number nr equal the number of nodes in
the function:
number of nodes in Rnr ,` (r) = nr

(4.2.10)

4.2.7 Orthogonality and nodes in the radial functions


Two 3D radial functions with the same ` but different energiesthat is, different radial quantum numbers
n0r 6= nr must be orthogonal (see Example 4.1):11
Z
hRn0r ,` |Rnr ,` ir
Rn0r ,` (r) Rnr ,` (r) r2 dr = 0,
if n0r 6= nr ,
(4.2.11)
|
{z
}
0
same `

where the subscript r on the Dirac bracket hRn0r ,` |Rnr ,` ir reminds us to integrate only over r, and Ive
assumed real radial functions (see the Aside on p. 271)
Rule: Radial functions with the same ` but different nr must be orthogonal.
10 Jargon: A node in a function f (r) is a positive value of r at which the function is zero, f (r) = 0. Even if f (r) = 0 at the
origin r = 0, that point is not considered a node.
11 Details: Equation (4.2.11) must hold because, for fixed `, the operator that acts on R
nr ,` (r) in the radial equationthe
thing in square brackets in Eq. (4.1.4), p. 266is Hermitian.

JQPMaster

Version: 8.35

Printed: August 17, 2010

272

4.2.7 Orthogonality and nodes in the radial functions

Orthogonality and radial nodes. The orthogonality requirement (4.2.11) explains why radial functions must have nodes. It also explains the positions of these nodes on the r > 0 axis. The radial function Rnr +1,` (r) must have one more node than Rnr ,` (r). This additional node must be located so Rnr +1,` (r)
will be orthogonal, a la Eq. (4.2.11), to all radial functions with radial quantum numbers between 0 and nr
(see Example 4.2):
Derivation of orthogonality condition. To understand the origin of this rule, we can reason by analogy from the orthogonality requirement for SAMEs nr ,`,m` (r, , ). This argument exploits the separable
form of a SAME, nr ,`,m` (r, , ) = Rnr ,` (r)Y`,m` (, ), and the orthonormality of the spherical harmonics:
I

Example 4.1 (Orthonormality of SAME radial functions.)


To begin, Ill denote the orthonormality integral by the rather graceless symbol
ZZZ
n r 0 ,`0 ,m` 0 (r) nr ,`,m` (r) d3 v
I(n0r , `0 , m0` | nr , `, m` )
Z

=
0

(4.2.12a)

n r 0 ,`0 ,m` 0 (r) nr ,`,m` (r) r2 dr sin d d,

(4.2.12b)

where I used the infinitesimal volume element in spherical coordinates, d3 v = r2 dr sin d d. Using
separability of nr ,`,m` (r, , ), I can (happily!) reduce the triple integral in Eq. (4.2.12b) to the product of a
purely radial integral times a purely angular integral whose value I know:
Z 2 Z
Z
Y`0 ,m0 (, )Y`,m` (, ) sin d d . (4.2.13a)
Rn0r ,`0 (r) Rnr ,` (r) r2 dr
I(n0r , `0 , m0` | nr , `, m` ) =
`
0
0
0
|
{z
}|
{z
}
hRn0r ,`0 | Rnr ,` ir
hY`0 ,m0` | Y`,m` i
,

In Dirac notation (Appendix G) the structure of Eq. (4.2.13a) emerges clearly:


I(n0r , `0 , m0` | nr , `, m` ) = hRnr ,` | Rnr ,` ir hY`0 ,m0` | Y`,m` i

(4.2.13b)

= hRn0r ,` | Rnr ,` ir `0 ,` m0` ,m` ,


In the second equality, I recognized that the angular integral equals the product of two delta functions for `
and m` (orthogonality of spherical harmonics). Because of `0 ,` , orthogonality of the functions { nr ,`,m` (r) }
does not require orthogonality of radial functions for different `; it requires orthogonality only of radial functions
with different radial quantum numbers nr and the same `.
We can interpret a SAME as a position probability density only if its radial function satisfies the radial
normalization condition
Z
hRnr ,` |Rnr ,` ir =

2
Rn
(r) r2 dr = 1
r ,`

normalization of radial functions

(4.2.14a)

Combining this condition with orthogonality [Eq. (4.2.11)], we get the radial orthonormality integral
Z
hRn0r ,` | Rnr ,` ir =

Rn0r ,` (r) Rnr ,` (r) r2 dr = n0r ,nr ,

(same `)

orthonormality
of radial functions.

(4.2.14b)

I Warning:

Dont forget that the integrand of an integral that contains radial functions R(r) must have
a factor of r2 . This factor comes from the three-dimensional volume element in spherical coordinates,
d3 v = r2 dr sin d d. Its easy to forget this factor. Dont.
I

JQPMaster

Example 4.2 (The orthogonality requirement and nodes in the radial function.)
The left panel in Fig. 4.2 shows R0,0 (r) and R1,0 (r) for a model of the three-electron lithium atom in its ground
state(Complement 4.D). For both states, ` = 0. The so-called 1s function R0,0 (r) has no nodes (nr = 0);
Version: 8.35

Printed: August 17, 2010

4.2.8 The essential degeneracy

273

the 2s function R1,0 (r) has one node (nr = 1). To illustrate how the orthogonality requirement determines
the position of this node, the right panel shows the integrand R1s (r)R2s (r)r2 of the radial orthonormality
integral (4.2.14b). The node in R1,0 (r) occurs at 0.802 a0 , precisely so as to ensure that the net shaded area
defined by this integrand equals zero.12

Orthogonality integrand Li 1s and 2s states

Radial functions: Li s states


4

0.10

0.05

Rn,{ HrL

u1,s HrL*u2,s HrL

1s
1
0

2s

-1
-2
0.0

0.5

1.0

0.00
-0.05
-0.10
-0.15

1.5

2.0

2.5

3.0

-0.20
0.0

r Ha0 L

0.5

1.0

1.5

2.0

2.5

3.0

r Ha0 L

Figure 4.2. Radial functions and the orthogonality integrand for s states of lithium. The minimumenergy (1s) radial function, R0,0 (r), has no nodes, while the radial function with the next highest energy, R1,0 (r) (the 2s function), has one node. This node is located so the contributions to the orthogonality
integral hR0,0 | R1,0 ir above and below the r axis are equal. The r 0 limits of the radial functions are
3/2
3/2
R1s (0) = 8.824 a0
and R2s (0) = 1.204 a0
, where a0 is the first Bohr radius 1 a0 ~2 /me e02 = 0.5292
A.

Try This!

4.4. Normalization here, normalization there.


Derive Eq. (4.2.14a) from the normalization condition on the SAME function n,`,m` (r).

4.2.8 The essential degeneracy


b n ,`,m (r) = En ,` n ,`,m (r)
The bound-state energies Enr ,` in the radial equation appear in the TISE H
r
r
r
`
`
b To each energy there corresponds one (or more) eigenfunctions n ,`,m (r). Because
as eigenvalues of H.
r
`
these eigenfunctions depend on m` , each SAME energy En,` may be degenerate.
This so-called essential degeneracy arises from the angular momentum degeneracy we discussed
in Chap. 3: each eigenvalue ~2 `(` + 1) of b
L 2 is (2`+1)fold degenerate, because for each ` there are (2`+1) linearly independent spherical harmonics Y`,m` (, ) for integral values of m` between m` = ` and m` = +`:13

Rnr ,` (r)

Enr ,`

nr ,`,` (r)

(r)

nr ,`,`1
..
.

nr ,`,`+1 (r)

nr ,`,` (r)

the essential degeneracy.

(4.2.15)

For example, corresponding to E2,1 (nr = 2, ` = 1), there are three linearly independent Hamiltonian
eigenfunctions, { 2,1,1 , 2,1,0 , 2,1,+1 }. Ergo, this energy is three-fold degenerate.
12 Commentary: The symbol a stands for the first Bohr radius, 1 a ~2 /m e 2 = 0.5292
A, the characteristic size of
e 0
0
0
atomic and molecular systems and the atomic unit of length (Appendix F).
13 Read on: You can learn more about degeneracies in central force problems from Degeneracies of the spherical well,
harmonic oscillator, and hydrogen atom in arbitrary degeneracies by Richard W. Shea and P. K. Aravind, American Journal
of Physics 64, 430, (1966).

JQPMaster

Version: 8.35

Printed: August 17, 2010

274

4.3.1 Bookkeeping, spectroscopic notation, and the principal quantum number

Rule: The degree of degeneracy of each bound-state SAME energy Enr ,` is at least (2` + 1):
g (Enr ,` ) (2` + 1)

essential degeneracy of a
bound-state SAME energy.

(4.2.16)

Notice that Eq. (4.2.16) is an inequality.14


Key Points
The structure of the 3D radial equation closely resembles that of the 2D radial equation. Moreover,
3D radial functions obey orthogonality conditions (and, for bound states, normalization conditions)
identical in form to those of 2D radial functions.
In 3D each value of ` spawns a different radial differential equation. Each such equation may have one
or more bound-state solutions, depending on the strength of the potential energy. Hence each energy
and radial function must, in general, carry two labels, ` and nr , the former to signify a particular radial
equation, the latter to signify a particular bound-state solution of that equation.
Orthogonality requirements impose a structure of nodes on all radial functions Rnr ,` (r) for a particular `: the number of nodes in Rnr ,` (r) equals the radial quantum number nr .
Each bound-state energy Enr ,` of a rotationally invariant system is at least (2` + 1)-fold degenerate.
The actual degree of degeneracy may be larger, depending on details of V (r).

4.3. The radial probability density


Whether we choose to credit the bizarre,
to take it seriously,
is finally irrelevant.
The world does its work.
July, July, by Tim OBrien

As in 1D and 2D, Hamiltonian eigenfunctions in 3D contain information about the probability of finding a
particle at various locations in space. For a SAME, the angular part of that information is in its spherical
harmonic Y`,m` (, ); the radial part is in its radial function Rnr ,` (r). Here well figure out how to most
insightfully extract that radial information. But first, we have to deal with a slight but significant change
in notation.
4.3.1 Bookkeeping, spectroscopic notation, and the principal quantum number
So far Ive labeled 3D SAMEs, radial functions, and energies by the radial quantum number nr , a nonnegative integer that distinguishes various solutions of the radial equation for a particular `. Largely for
historical reasons, its conventional to use instead the principal quantum number n, an integer that, for
a particular `, has a minimum value ` + 1. So instead of Rnr ,` (r), we write Rn,` (r).15
Although little more than bookkeeping, this change in labeling can confuse you if youre unfamiliar with
it. The key is the relationship between the radial quantum number nr and the principal quantum number n
for a particular `. Since nr = 0 corresponds to n = nmin = ` + 1, we have
14 Commentary: The degree of degeneracy of an arbitrary SAME of a one-electron atom, for example, is greater than (2` + 1).
Ignoring spin, and treating the atom in the pure-Coluomb approximation (Chap. 5) gives a degree of degeneracy n2 , where n
is the so-called principal quantum number n = nr + ` + 1. Taking into account electron spin (Chap. 6) increases this degree
of degeneracy to 2n2 .
15 A cautionary note: This convention is not applied to all rotationally invariant systems. An important exception is the 3D
isotropic harmonic oscillator (footnote 2). For convenience, SAMEs and radial functions for this system are labeled by nr ,
although (confusingly) this index is called the principal quantum number and is denoted by n. In any case, for this system the
ground-state radial function is labeled by n = 0 and has no nodes.

JQPMaster

Version: 8.35

Printed: August 17, 2010

4.3.1 Bookkeeping, spectroscopic notation, and the principal quantum number


nr = n ` 1

275

the radial and principal


quantum numbers (3D).

n = nr + ` + 1

(4.3.1)

To illustrate, Ive tabulated various notations for a few radial functions in Tbl. 4.1. Ive also included the
so-called spectroscopic notation for each state, a symbology you may have encountered previously; this
notation is summarized in Tbl. 4.2.16
Key point! Accompanying this change in convention is a change in the rule in 4.2.6 about nodes in radial
functions. This rule now reads: To ensure orthogonality of radial functions with different principle quantum
numbers n but the same `, each radial function Rn,` (r) must have nr nodes, where
number of nodes in Rn,` (r) = nr = n ` 1

(4.3.2)

These nodes must be positioned along the r axis, in such a way that the orthogonality condition Eq. (4.2.14b)
will be satisfied for all radial functions Rn,` (r) for that `that is, for n = ` + 1, ` + 2, . . . .
Rnr ,` (r)

Rn,` (r)

spectroscopic notation

number of nodes

R0,0 (r)
R1,0 (r)
R0,1 (r)
R2,0 (r)

R1,0 (r)
R2,0 (r)
R2,1 (r)
R3,0 (r)

1s
2s
2p
3s

0
1
0
2

Table 4.1. Radial functions labeled by the radial quantum number and the
principal quantum number. The third column shows conventional spectroscopic notation (Tbl. 4.2), while the fourth shows the number of radial nodes in Rn,` (r), which
is nr = n ` 1.

Table 4.2. Spectroscopic notation. For states with ` > 3, the


spectroscopic designations increase
alphabetically, so ` = 4 states are
labeled g states and so forth.
The names in the last column are
of historical interest only; they refer to particular spectral series observed when atoms undergo transitions (Chap. 18) and are the origin
of the letters s, p, d, and f .

Try This!

notation

degree of degeneracy

s state
p state
d state
f state

0
1
2
3

0
3
5
7

historical name
sharp
principal
diffuse
fundamental

4.5. Probably too much practice with notation for radial functions.
Extend Tbl. 4.1 to include functions with nr = 2 and ` = 1. Also include rows for n = 3, ` = 1 and
n = 3, ` = 2. Finally include rows for states with spectroscopic notation 4s, 4p, 4d, and 4f . (If youre
still unsure that you understand how to deal with this change of notation, continue developing exercises
like this until you become confident or just cant stand it anymore.)

16 Commentary: Spectroscopists devised this notation in the early days of quantum physics. By convention, the spectroscopic
notation for a bound state with principal quantum number n, orbital angular momentum quantum number `, and projection
quantum number m` is n<letter>m` , where <letter> is s, p, d, f, g, . . . for ` = 0, 1, 2, 3, 4, . . . . Well use this notation
throughout our discussion of atoms, beginning in Chap. 5. (If you think this convention is arbitrary, wait till you see spectroscopic
notation for molecular states.)

JQPMaster

Version: 8.35

Printed: August 17, 2010

276

Try This!

4.3.3 Derivation and interpretation of the radial probability density


4.6. The usefulness of convention.
Suppose we had defined the principal quantum number n in such a way that, for each set { un,` (r) }, the
state of lowest energy was labeled n = 1, the next lowest n = 2, and so forth. List the number of nodes
in the first five functions un,` (r) for ` = 0, 1, 2, and 3. This should convince you that the admittedly
less intuitive convention nmin = ` + 1 leads to results that are easier to remember.

4.3.2 The position probability density according to Born


We ascribe physical significance to a wave function (r, t) via the Born interpretation (Appendix A):

Rule: The probability of finding a particle in state (r, t) in an infinitesimal volume element d3 v at (vector)
position r is the product of the position probability density P(r, t) times d3 v, where
P(r, t) |(r, t)|2 = (r, t) (r, t)

position probability density.

(4.3.3)

If the state is stationary, so (r, t) = E (r) ei Et/~ , then the 3D probability density (4.3.3) looses its time
dependence and becomes

(r) E (r)
PE (r) |E (r)|2 = E

(for a stationary state).

(4.3.4)

For a bound-state SAME, with E (r) = n,`,m` (r), the probability density (4.3.4) looses its dependence
and becomes

Pn,`,m` (r, ) = n,`,m


(r)n,`,m` (r)
`

(4.3.5a)

= Rn,`
(r) Y`,m
(, )Y`,m` (, ), by the separable form of n,`,m` (r),
`

1 2
R (r) |`,m` ()|2 ,
2 n,`

(4.3.5b)

1
using Y`,m` (, ) = `,m` () ei m` .
2

(4.3.5c)

4.3.3 Derivation and interpretation of the radial probability density


The next step in simplifying the probability density is to change the question. Instead of asking for the
probability of finding a particle in d3 v at a vector position r, we ask for the probability of finding the
particle in a spherical shell of thickness dr at radial position r. As you can see in Fig. 4.1, the answer
requires that we sum the probability density over all angles:
Rule: The probability of finding a particle with stationary-state wave function E (r) in a spherical shell of
thickness dr at a distance r from the origin is17
Z 2 Z
Z 2 Z

PE (r, ) d3 v =
E
(r) E (r) r2 dr sin d d .
(4.3.6)
| {z }
0
0
0
0

I Warning:

Even though we are integrating (summing) over all angles, as indicated by the underbrace in Eq. (4.3.6), the resulting integral must include the factor r2 dr from the 3D volume element
d3 v = r2 dr sin d d.18
17 Details:

Notice that we do not integrate over r. Thats because we want the probability density at a particular value of r.
This factor differs from the factor of that arises from the 2D volume element d2 v in the 2D radial probability

18 Commentary:

density.

JQPMaster

Version: 8.35

Printed: August 17, 2010

4.3.3 Derivation and interpretation of the radial probability density

277

Figure 4.1.
A spherical shell
used to define the radial probability density. The shell is located
at radius r, and its thickness is dr.
The radial probability density evaluated at r times dr is the probability
of finding a particle in this shell, irrespective of its angular position.

Deriving the radial probability density of a SAME. For a SAME, the separable form Eq. (4.3.5b)
allows us to perform the angular integrals in Eq. (4.3.6) analytically:
Z
0

Z
PE (r, ) d3 v =

Rn,`
(r) Y`,m
(, )Y`,m` (, ) r2 dr sin d d,
`

2
Rn,`
(r) r2

2
0

Y`,m
(, )Y`,m` (, ) sin d d,
`

2
= Rn,`
(r) r2 ,

(4.3.7a)
(4.3.7b)
(4.3.7c)

where the last step follows from orthonormality of the spherical harmonics. We have found a simpler version
of the Rule in 4.3.2:
Rule: The probability of finding a particle in a SAME n,`,m` (r) in a spherical shell of thickness dr at r is
rad
Pn,`
(r) dr, where the radial probability density is
rad
2
Pn,`
(r) r2 Rn,`
(r)

radial probability density


for a 3D SAME

(4.3.8)

rad
(r) are
Understanding the radial probability density of a SAME. Radial probability densities Pn,`
essential interpretive tools for central-force systems, especially atoms. Interpreting these functions correctly
rad
requires an appreciation of what they do not tell us. Neither Pn,`
(r) nor any quantity calculated from it
contains information about where inside the shell the particle might be: this angular information is washed
out by the angular integration in Eqs. (4.3.7).

The radial probability density and normalization of the radial function. According to our interrad
pretation of Pn,`
(r), if we add radial probability densities for shells of thickness dr at all possible radii,
0 r < , we must get unityafter all, the particle has to show up somewhere:
Z
Z
rad
2
(4.3.9)
Pn,`
(r) dr =
Rn,`
(r) r2 dr = 1.
0

Equation (4.3.9) is consistent with and equivalent to the normalization condition for Rn,` (r) [Eq. (4.2.14a),
p. 272].
Try This!

4.7. Orthonormality in action!


The particular mathematical structure of a SAME n,`,m` (r) lets us perform the angular integrations
in Eq. (4.3.6) without doing any algebra. Can we perform this legerdemain for an arbitrary wave function (r, t) of a rotationally invariant system?

JQPMaster

Version: 8.35

Printed: August 17, 2010

278

4.3.3 Derivation and interpretation of the radial probability density

Example 4.3 (Radial probability densities for the ground state of atomic lithium.)
Plots of radial probability densities for an atom are cornucopias of insights. For instance, Fig. 4.2 shows radial
probability densities for the 1s and 2s states of the model lithium atom of Complement 4.D [see also Fig. 4.2,
p. 273].

Noteworthy features of Fig. 4.2:


(1) In the probabilistic sense typical of quantum mechanics, an electron in a 1s state is more localized near
the origin than one in a 2s state. (This observation exemplifies the shell model of the atom, which will
assume center stage in Part III.)
rad
(2) The maximum probability of finding a 2s electron, the peak of P2s
(r), is much smaller than that of
rad
rad
finding a 1s electron, the peak of P1s (r). [The peak value of P2s (r) is 0.267 at r = 3.10 a0 , while the
rad
peak value of P1s
(r) is 1.456 at r = 0.372 a0 .]

(3) There is a tiny but nonzero probability of finding a 2s electron very near the origin, at r = 0.305 a0
J
(where the magnitude is 0.023). This little peak is due to the node in R2s (r) of Fig. 4.2 (4.2.7).

Radial probability densities: Li s states


1.5

Figure 4.2. Radial probability densities


for 1s and 2s electrons in atomic lithium.
These radial probability densities were calculated as per Eq. (4.3.8) from the radial
functions in Fig. 4.2, p. 273. They are
based on the model in Complement 4.D.
The units of radius are bohr ( a0 ), where
1 a0 ~2 /me e02 = 0.5292
A. The radial probability densities are in units of a3
0 , as required to make the integrand in Eq. (4.3.9)
dimensionless.

1s

1.0
rad
HrL
Pn,{

0.5

0.0

2s
0

r Ha0 L

Example 4.4 (Radial probability densities for 1s states of several atoms.)


Plots of radial probability densities also facilitate comparison of properties of different atoms. Figure 4.3
rad
compares Pn,`
(r) for 1s electrons (n = 1 and ` = 0) in several atoms: lithium, (Li: Z = 3); boron, (B: Z = 5);
nitrogen, (N: Z = 7); and neon, (Ne: Z = 10).19
Noteworthy features of Fig. 4.3:
rad
(1) The radius of the peak in P1s
(r) decreases with increasing atomic number Zas does the average
radius where the electron would be found, from hri1s = 0.558 a0 for Li to hri1s = 0.155 a0 for Ne (4.3.5).
rad
rad
Since the integral under P1s
(r) must equal 1 for any atom, the magnitude of the peak in P1s
(r) must
increase with increasing Z.

(2) The spatial extent of the radial functionsthe range of r over which the radial probability density
is appreciabledecreases with increasing Z. Physically, therefore, we are more uncertain about the
position of a 1s electron in Li than in, say, Ne: the radial uncertainty for Li is (r)1s = 0.322 a0 ; while
J
for Ne, its (r)1s = 0.090 a0 (4.3.5).
19 Jargon: The atomic number Z of an atom equals the number of protons in its nucleus. For a neutral atom, Z also equals
equals the number of bound electrons.

JQPMaster

Version: 8.35

Printed: August 17, 2010

4.3.4 Integrated probability densities: (finite) shell and sphere probabilities

279

1s radial probability densities

Figure 4.3. Radial probability densities


for 1s electrons in several atoms. These
radial probability densities were calculated according to Eq. (4.3.8) from the radial functions of Morse et al. (1935). As in Fig. 4.2,
the units of radius are bohr ( a0 ), where
1 a0 ~2 /me e02 = 0.5292
A. The radial probability densities are in units of a3
0 .

radial probability density

6
5

Ne
4

2
1
0
0.0

Li
0.2

0.4

0.6

0.8

1.0

1.2

1.4

rHa0 L

4.3.4 Integrated probability densities: (finite) shell and sphere probabilities


No experiment can actually measure a particles position in an infinitesimal region of space. The closest
measurable quantity to the radial probability density is the probability of finding the particle anywhere in
a very thin spherical shell, of finite thickness r, at some radius r0 . Heres a sensible way to define this
quantity, which Ill call a shell probability ,
Z
shell
Pn,`
(r0 )

r0 + 21 r

r0 12 r

Z
rad
Pn,`
(r) dr

r0 + 12 r

r0 12 r

2
Rn,`
(r) r2 dr,

shell probability at r0 .

(4.3.10a)

This shell lies between two spheres, one of radius r0 r/2, the other of radius r0 + r/2 (Fig. 4.1, p. 277).
A related quantity is the probability of finding the particle anywhere within a sphere of radius r0 . The
corresponding sphere probability is the sum (integral) of shell probabilities for all shells with radii
0 r r0 :
Z r0
Z r0
sphere
rad
2
Pn,` (r0 )
Pn,` (r) dr =
(4.3.10b)
Rn,`
(r) r2 dr,
sphere probability at r0 .
0

rad
Like any quantity based on Pn,`
(r), the sphere probability contains no angular information. Moreover, it
tells us nothing probability of finding the particle at a particular radius inside the sphere.

JQPMaster

Example 4.5 (The sphere probability for electrons in the ground state of atomic lithium.)
sphere
In Fig. 4.4, I show sphere probabilities Pn,`
(r0 ) for 1s and 2s electrons in the ground state of lithium.
sphere
sphere
Unsurprisingly, Pn,` (r0 ) goes to zero as r 0. Equally unsurprisingly, Pn,`
(r0 ) goes to 1 as r . [In
that limit, the definition Eq. (4.3.10b) reduces to the normalization integral, Eq. (4.2.14a), p. 272]. Plots of
sphere
rad
sphere probabilities Pn,`
(r0 ) offer insights hard to see in plots of Pn,`
(r) like those in Fig. 4.2, p. 278. For
sphere
example, the radius where Pn,` (r0 ) 1 is a reasonable measure of the size of the electrons state. For
the ground state of lithium, the 1s electron, the sphere probability is 0.999 by r0 = 2.1 a0 . For the much less
localized 2s electron, the sphere probability attains this value at r0 = 12.0 a0 .

Version: 8.35

Printed: August 17, 2010

280

4.3.5 Radial expectation values and uncertainties


Sphere probability for Li
1.0

Figure 4.4.
Sphere probability for 1s
and 2s electrons in the ground state of
lithium. The horizontal axis shows the radius r0 in the definition Eq. (4.3.10b). The
thin horizontal line at 0.5 on the vertical axis
separates spheres that contain less than 50%
of the position probability (below the line)
from those that contain more than 50% (above
the line). These calculations were based on
the radial functions of Morse et al. (1935).
The sphere radii are in units of bohr ( a0 ), and
the sphere probabilities are dimensionless.

sphere probability

0.8

1s

0.6

2s

0.4
0.2
0.0

10

sphere radius Ha0 L

4.3.5 Radial expectation values and uncertainties


sphere
rad
shell
Plots of the radial probability density Pn,`
(r) and finite probabilities like Pn,`
(r0 ) and Pn,`
(r0 ) show
the radial distribution of the position probability of a particle. In many cases, however, we want one or two
numbers that convey the same type of information, albeit in less detail. The most useful such numbers are
the expectation value and uncertainty in the radial position of the particle.

The radial expectation value for a SAME. For a completely general state (r, t), expectation values
and uncertainties depend on time, and in 3D, their evaluation requires doing triple integrals (ugh). The
radial expectation value, for example, is
Z
hri(t) h(t) | r | (t)i = (r, t) r (r, t) d3 v
Z

(4.3.11)

(r, t) r (r, t) r dr sin d d.


0

For a stationary state, the time dependence goes away, as usual, but were still confronted by a triple integral:
Z 2 Z Z

(4.3.12)
hriE = hE | r | E i =
E
(r) r E (r) r2 dr sin d d.
0

For a SAME, however, the separable structure n,`,m` (r, , ) = Rn,` (r)Y`,m` (, ) allows us to perform the
angular integration analytically. As in our derivation of the radial probability density [Eqs. (4.3.7), p. 277],
all that remains is a single radial integral :
Z
hrin,` hn, `, m` | r | n, `, m` i =
Rn,` (r) rRn,` (r) r2 dr
(4.3.13a)
| {z }
0
Z
Z
rad
2
r Pn,`
(r) dr.
(4.3.13b)
Rn,`
(r) r3 dr =
=
0

As the subscripts on hrin,` indicate, the mean radius depends on the principal and orbital angular momentum
quantum numbers but not on m` .
Now, look carefully at each step in Eqs. (4.3.13). In Eq. (4.3.13a), the underbrace emphasizes the
contribution r2 dr, all that remains of d3 v = r2 dr sin d d after angular integration. This contribution
contains the familiar factor of r2 . Thats why the power of r in the integrand in Eq. (4.3.13b) is r3 , not r.
rad
2
The second equality in Eq. (4.3.13b) emphasizes the role of Pn,`
(r) = r2 Rn,`
(r):
rad
Rule: The radial probability density Pn,`
(r) acts as a weighting factor for each value of r where the particle

might be found. So the average radial position hrin,` , is, sensibly, the weighted mean of all possible radii.

JQPMaster

Version: 8.35

Printed: August 17, 2010

281
The radial uncertainty for a SAME.

Similar machinations yield the radial uncertainty:


q
(r)n,` = hr2 in,` hri2n,` ,

where hr2 in,` = hn, `, m` | r2 | n, `, m` i


Z
2
=
Rn,`
(r) r4 dr.

(4.3.14a)
(4.3.14b)
(4.3.14c)

I used Eqs. (4.3.13b) and (4.3.14) to calculate the results in Examples 4.3 and 4.4.

I Warning:

The one-variable integrals in Eqs. (4.3.13b) and (4.3.14) pertain only to SAMEs. For an
arbitrary stationary state, you must evaluate three-variable integrals as in Eq. (4.3.12). For an arbitrary
non-stationary state, this integral will depend on time, as in Eq. (4.3.11).

Try This!

4.8. Labels galore.


Why does hri for a SAME depend on n and ` but not on m` ? For a potential energy V (r, ) that depends
on as well as r, on which of the quantum numbers n, `, m` will hri depend? Explain your answer.

4.4. The case of the disappearing derivative: The reduced radial equation
One of Sherlock Holmess defectsif, indeed, one may call it
a defectwas that he was exceedingly loath to communicate his full plans
to any other person until the instant of their fulfillment.
The Hound of the Baskervilles, by Sir Arthur Conan Doyle

The radial equation

2
b rad (r) + ~ `(` + 1) + V (r) RE,` (r) = E RE,` (r),
T
2mr2

` = 0, 1, 2, . . .

(4.4.1a)

poses a couple of practical challenges. The first challenge is that its a differential equation. The second
comes from the radial kinetic-energy operator,

~2 1 d
~2
d
2 d
radial kinetic-energy
rad
2 d
b
(4.4.1b)
T (r) =
r
=
+
.
2
2
operator
2m r dr
dr
2m dr
r dr
|
{z
}
b rad renders Eq. (4.4.1a) a second-order ordinary differential equation with
As the underbrace emphasizes, T
a first-derivative term:

~2
d
2 d
~2 `(` + 1)

+
+
+
V
(r)
RE,` (r) = E RE,` (r),
` = 0, 1, 2, . . .
(4.4.1c)
2m dr2
r dr
2mr2
The presence of the first-derivative, (2/r) dR(r)/dr, makes solving this equation more difficult than it needs
to be. Moreover, it makes the mathematical structure of the 3D radial equation different from that of the
1D TISEwhich precludes our adapting insights from the study of 1D quantum mechanics to the solution
of this equation.
4.4.1 Introducing the reduced radial equation
We can make these problems go away by getting rid of the first derivative term. Of course, I cant just
erase the first derivative term because I dont like it. What I can do is define a new functionthe reduced
radial function uE,` (r)that satisfies a second-order differential equation which has no first derivative
term and from which I can determine RE,` (r). There exists such a function uE,` (r), and its very simply
JQPMaster

Version: 8.35

Printed: August 17, 2010

282

4.4.2 Boundary conditions on the reduced radial function

related to RE,` (r):


un,` (r) r Rn,` (r) = Rn,` (r) =

1
un,` (r)
r

definition of the reduced radial function.

(4.4.2)

Plugging Eq. (4.4.2) into Eq. (4.4.1c) and expanding out the derivatives leads to an equation for uE,` (r),
the reduced radial equation

~2 d2
~2 `(` + 1)

+
+ V (r) En,` un,` (r) = 0
2m dr2
2mr2

reduced radial equation.

(4.4.4)

In terms of the reduced radial function un,` (r), the separable form a SAME becomes
n,`,m` (r, , ) =

1
un,` (r) Y`,m` (, )
r

(4.4.5)

Whoa! Stop! Right now a question should be bothering youat least, this question bothered me when I first
encountered this step. How did he know that the definition in Eq. (4.4.2) would eliminate the first derivative
from Eq. (4.4.1c)more to the point, how am I supposed to figure out how to perform this trick when Im confronted
with a similar situation? There is, it turns out, a systematic procedurea cookbook for eliminating unwanted
derivative terms from a differential equation . You can apply this strategy to any differential equation involving
any variable or derivative. If youre interested, Ive developed and applied this strategy to the radial equation
in Example 4.12 in Complement 4.A.

4.4.2 Boundary conditions on the reduced radial function


Immediately upon deriving a differential equation, we should determine the boundary conditions its solutions
must obey. The boundary conditions on un,` (r) follow directly from the conditions on the radial functions
Rn,` (r) = un,` (r)/r (4.2.5). For bound states, [Eq. (4.2.8), p. 270],20
Rn,` (0) = a finite number

un,` (0) = 0,

(4.4.6a)

Rn,` (r) 0

un,` (r) 0.

(4.4.6b)

To ensure normalizability of the radial function Rn,` (r) we need only require that its value at the origin be a
finite number. But the reduced radial function un,` (r) is the radial function multiplied by 1/r, and 1/r blows
up as r 0. So un,` (r) must go to 0 as r 0: thats the point boundary condition in Eq. (4.4.6a). The
limit boundary condition in (4.4.6b) for un,` (r) is the same as for Rn,` (r).

Example 4.6 (Reduced radial functions for the ground state of atomic lithium.)
To show you the effect of the modest little factor r in the definition Eq. (4.4.2), Ive plotted in Fig. 4.1
the functions u1s (r) and u2s (r) that correspond to the radial functions R1s (r) and R2s (r) in the left panel
in Fig. 4.1. The radial probability densities for these states are in Fig. 4.2, p. 278 and discussed in Example 4.3.
The contrasts between Rn,` (r) and un,` (r) are striking. The boundary condition that u1s (0) = 0 forces
this function to turn over as r 0, resulting in a peak at r = 0.372 a0 thats not present in R1s (r). Similarly, u2s (r) has a minimum at r = 0.305 a0 and a maximum at r = 3.099 a0 . These features appear in the

20 Details: Although the boundary condition u


n,` (0) = 0 properly identifies most physically admissible radial functions, it
is more stringent than absolutely necessary. The true origin of the boundary condition at the origin is the requirement that
b rad + T
b ang + V (r) be Hermitian; this must be satisfied lest the eigenfunctions of this operator be physically
the Hamiltonian T
inadmissible. As discussed in 12.4 of Merzbacher (1998), this Hermiticity requirement leads to a less stringent boundary
conditions at the origin which, in most cases, can be replaced by Eq. (4.4.6a). See also Chap. 6 of Galindo and Pascual (1991).

JQPMaster

Version: 8.35

Printed: August 17, 2010

4.4.3 Subduing the reduced radial equation through pattern matching


Radial functions: Li s states

Reduced radial Function: Li s states

Rn,{ HrL

283

1.2

1.0

0.8

1s
1

un,{ HrL

1s

0.6

2s

0.4
0

-2
0.0

0.2

2s

-1
0.5

1.0

0.0
1.5

2.0

2.5

3.0

-0.2

r Ha0 L

10

r Ha0 L

Figure 4.1. Radial and reduced radial functions for 1s and 2s states of atomic lithium. The
reduced radial functions in the right panel were were calculated from the radial functions in the left panel
[see Fig. 4.2, p. 273]. Both are based on the radial functions of Morse et al. (1935). The units of radius
1/2
are bohr ( a0 ), where 1 a0 ~2 /me e02 = 0.5292
A. The units of the reduced radial functions are a0
.

radial probability densities, in Fig. 4.2, p. 278 and hence influence quantities like the sphere probabilities of Example 4.5. In fact, its much easier to predict and understand the radial distribution of position probability
density from reduced radial functions than from the radial functions themselves.
J

4.4.3 Subduing the reduced radial equation through pattern matching


4.4.3.1. The structure of the reduced radial equation
Ive already introduced The Principle of Pattern Matching : Pursue familiar patterns (Appendix Q).
The key to pattern matching is mindful scrutiny: you must keep your brain engaged, not just your optic
nerve. Look again at the reduced radial equation, this time focusing on its mathematical structure:
(
2
)
~2 d2
~ `(` + 1)

(4.4.7)
+
+ V (r) un,` (r) = En,` un,` (r).
2m dr2
2mr2
|
{z
}
function of r

We see a second-derivative of the unknown function un,` (r). Added to this is a function of r. One term in this
function, the barrier term came from action of the angular kinetic-energy operator on a spherical harmonic;
the other is the physical potential energy V (r), a consequence of whatever forces act on the particle. On the
right-hand side, we see the bound-state energy En,` times the unknown function.21
21 Jargon: The barrier term is usually called the centrifugal potential energy . This term is somewhat misleading, because,
as weve seen, the term ~2 `(` + 1)/2mr2 comes from action of the angular kinetic-energy operator. This term centrifugal
potential energy comes from classical physics. Until about the last 2/3 of the 20th , textbooks in classical mechanics, when
talking about central-force systems, introduced a fictitious force they called the centrifugal force. (Nowadays, this force,
which is still fictitious, is usually called the centripetal force.) Whateve you call it, the classical potential energy that
corresponds to this fictitious force is L2 /2mr2 . In the quantum mechanical radial and reduced radial equations for a rotationally
invariant system, the centrifugal potential energy looks just like the classical expression L2 /2mr2 with the classical orbital
angular momentum L2 replaced by its eigenvalue in the SAME for these equations, ~2 `(` + 1). But its actually an angularkinetic-energy term.

JQPMaster

Version: 8.35

Printed: August 17, 2010

284

4.4.4 Orthonormality of the reduced radial function

4.4.3.2. The effective potential energy


Now, where have you seen an equation like Eq. (4.6.3) before? Right: in your study of 1D quantum mechanics.
The 1D TISE for a bound-state wave function E (x) of a particle of mass m with potential energy V (x) has
precisely the same structure as Eq. (4.6.3):

~2 d2
n (x) = En n (x),
TISE (1D).

+
V
(x)
(4.4.8)
2m dx2
The role of V (x) in the 1D TISE is played by the underbraced term in Eq. (4.6.3). This term is not the
physical potential energy: the physical potential energy is V (r). But structurally, this term plays the role of
a potential energy in the reduced radial equation. So Ill call it the effective potential energy :
V`eff (r)

~2 `(` + 1)
+ V (r)
2mr2

the effective potential energy.

(4.4.9)

With this definition, we can make the reduced radial equation look even more like the 1D TISE:

~2 d2
eff

+ V` (r) un,` (r) = En,` un,` (r)


2m dr2

reduced radial equation in terms


of the effective potential energy.

(4.4.10)

I Warning: Dont get carried away with the similarities between the reduced radial equation Eq. (4.4.10)
and the 1D TISE Eq. (4.4.8). Two differences are paramount and should be memorized:
(1) The effective potential energy V`eff (r) in the reduced radial equation is not the physical potential energy.
Rather, V`eff (r) is an artifice that appeared as a result of our derivation of this equation.
(2) The domain of the independent variable r in the reduced radial equation, 0 r < , is not the same
as the domain of x in the 1D TISE, < x < .22
Key point! The physical potential energy is the potential energy of the system. It appears alongside the
b + V (r). The physical potential energy represents whatever
b=T
kinetic-energy operator in the Hamiltonian H
actual forces act on the particle. Ultimately, the physical potential energy determines the properties of the system,
including the total number of bound states.

4.4.4 Orthonormality of the reduced radial function


In addition to satisfying the reduced radial equation (4.4.4) and the boundary conditions (4.4.6), un,` (r) satisfies a bunch of useful mathematical properties that follow from its definition un,` (r) r Rn,` (r). Foremost
among these is its orthonormality integral . As we learned in 4.2, the radial functions obey [Eq. (4.2.14b),
p. 272]
Z

hRn0 ,` | Rn,` ir =

Rn0 ,` (r)Rn,` (r) r2 dr = n0 ,n .

(4.4.11a)

22 Commentary: Although noting the domain may seem like a mathematical fine point, it has significant consequences for
determining and interpreting the reduced radial function. Some authors emphasize this point by defining the effective potential
energy (4.4.9) as
2
~ `(` + 1)
+ V (r)
r0
eff
2
V` (r)
2mr
0
r<0

While theres nothing wrong with this definition, I prefer Eq. (4.4.9), since, by definition,the radial variable r cannot, be
negative.

JQPMaster

Version: 8.35

Printed: August 17, 2010

4.4.5 Summary: Why bother with the reduced radial function?

285

Inserting Rn,` (r) = un,` (r)/r into this equation yields the orthonormality relation for the reduced radial
function:
Z
hun0 ,` | un,` ir =

un0 ,` (r)un,` (r) dr = n0 ,n


0

orthonormality of bound-state
reduced radial functions.

(4.4.11b)

Notice that the factor r2 , which has been present in every radial integral so far, is gone. It was canceled by
the factor 1/r2 from the definition Rn,` (r) un,` (r)/r.

I Warning:

Remember that Eqs. (4.4.11) apply only to radial functions with the same orbital quantum
number `. No such relation pertain to functions with `0 6= `, whether or not the principal quantum numbers n0
and n in the functions are equal.

4.4.5 Summary: Why bother with the reduced radial function?


The structural similarity between the 1D TISE (4.4.8) and the reduced radial equation (4.4.10) allows us to
adapt our understanding of bound-states in 1D to 3D rotationally invariant system:23
Skills for drawing qualitative sketches of the eigenfunctions n (x) [9.3].
Insights into the influence of symmetry properties on the eigenfunctions n (x) [9.3].
Known solutions for special cases such as the 1D infinite square well [7.4], the finite square well [8.8],
and the simple harmonic oscillator [9.8].
Powerful problem-solving techniques such as power-series solution of differential equations [9.7], separation of variables [7.1], and numerical methods for solving transcendental [8.8] and differential
equations.
Insights into the structure of Hamiltonian eigenfunctions and eigenvalues for a generic potential energy, such as the nature of bound and continuum states, the effect of nodal structure on eigenfunctions,
and so forth.
To show you how to perform manipulations using un,` (r), Ive included a non-trivial example, Example 4.13,
in Complement 4.A.
Key Points
Introduction of the reduced radial function un,` (r) r Rn,` (r) eliminates the first-derivative term
from the system-dependent differential equation we must solve to complete specification of a SAME
[Eq. (4.4.5), p. 282].
The differential equation for the reduced radial function, Eq. (4.4.4), p. 282, contains, in addition to the
physical potential energy, V (r), a term ~2 `(` + 1)/2mr2 that arises from action of the angular kineticenergy operator on a spherical harmonic in the separable form n,`,m` (r, , ) = Rn,` (r)Y`,m` (, ). Because the sum of these two terms, the effective potential energy V`eff (r) ~2 `(` + 1)/2mr2 + V (r),
appears in the reduced radial equation, this sum governs the mathematical behavior of the un,` (r).
The reduced radial functions satisfy boundary conditions at the origin and in the asymptotic limit
[Eq. (4.4.6), p. 282] that follow from the boundary conditions on the radial function. They key difference
is that at the origin, Rn,` (r) must equal to a constant, while un,` (r) must equal zero.
Three-dimensional integrals written in terms of reduced radial functions un,` (r) will not contain the
additional factor of r2 that must be present when these integrals are written in terms of radial functions Rn,` (r).
23 Read on: In this list, the parenthetical sections refer to Understanding Quantum Physics (UQP), which, of course, I
urge you to rush out and purchase.

JQPMaster

Version: 8.35

Printed: August 17, 2010

286

4.5.1 The competition between the barrier term and the physical potential energy

Aside. The radial momentum.


It is often useful in problem solving to write the 3D radial kinetic-energy operator terms of the radial
momentum operator , which is defined as

b
pr i ~

1
+
r
r

1
= i ~
r

r
r

radial momentum operator.

(4.4.12a)

In terms of b
pr , the radial kinetic energy assumes the familiar form24
p2r
b rad = b
T
2m

(4.4.12b)

Youll have a chance to properties of this operator in Problem 4.10.


Try This!

4.9. The radial kinetic-energy and radial-momentum operators.


Show (or otherwise convince yourself) that plugging Eq. (5.C.4) for the radial momentum
b rad gives Eq. (4.4.1b) for the radial kinetic-energy operator.
operator b
pr into Eq. (4.4.12b) for T

4.5. The roles of the physical and effective potential energies


It is of the highest importance in the art of detection to be able to recognize, out of a
number of facts, which are incidental and which are vital. Otherwise your energy and
attention must be dissipated instead of being concentrated.
Sherlock Holmes, in The Reigate Squires, by Sir Arthur Conan Doyle

The effective potential energy for a spherically symmetric physical potential energy, [Eq. (4.4.9), p. 284]
V`eff (r) V (r) +

~2 `(` + 1)
,
2mr2

(4.5.1)

plays a vital role in bound states of a 3D rotationally invariant system. Indeed, it holds the key to the
qualitative behavior of the reduced radial function un,` (r).
4.5.1 The competition between the barrier term and the physical potential energy
The definition Eq. (4.5.1) of the effective potential energy reveals two key features of the barrier term
~2 `(` + 1)/2mr2 :
(1) The barrier term is positive for all ` > 0 but equals zero if ` = 0.
(2) The barrier term is proportional to 1/r2 , and therefore goes to zero as r increases from r = 1 to r .
As r decreases from r = 1 to zero, however, this term grows very rapidly, and in the limit r 0, it
goes to +.25
In the next example, well examine the contributions to the effective potential energies for a typical system.

Example 4.7 (The effective potential energy for s-state electrons in the ground state of lithium.)
Figures 4.1 illustrate the effects of the barrier term on the effective potential energy for the model for the
lithium in Complement 4.D. Scrutinize this figure to see how it illustrates the following rule:
Rule: For states with ` > 0, the effective potential energy is determined by a competition between the physical
potential energy V (r), which must be attractive enough to support at least one bound state, and the barrier

24 Read on: For mathematical reasons b


pr cannot represent an observable (it lacks self-adjoint extensionssee 6.2 of Galindo
and Pascual (1991)). The square of this operator, however has no such deficiency and so can represent an observable.
25 Details: For most potential energies in nature, the asymptotic decrease to zero occurs as 1/r p for a constant p 1.

JQPMaster

Version: 8.35

Printed: August 17, 2010

4.5.2 The effective potential energy and the number of bound states

287

term ~2 `(` + 1)/2mr2 , which for ` > 0 is repulsive (positive) for all r. The magnitude of this term increases rapidly
as r 0.

{=3
V{ eff HrL

{H{+1L2r 2

effective potential energy

potential energy

-2

{=1

{=2

-2

{=0

VHrL

-4
0.0

0.5

1.0

1.5

2.0

2.5

-4
0.0

0.5

r Ha0 L

1.0

1.5

2.0

2.5

r Ha0 L

Figure 4.1. Effects of the centrifugal barrier on the effective potential energy. Left panel: A physical
potential energy V (r), the centrifugal barrier term ~2 `(` + 1)/2mr2 , and the effective potential energy V`eff (r)
for the model for lithium in Complement 4.D. The orbital angular momentum quantum number is ` = 1. Right
panel: The effective potential energy for ` = 0, 1, 2, and 3.
The left panel in Fig. 4.1 shows this competition in action. The physical potential energy V (r) is purely
attractive. As r increases, the barrier term decreases: by about r & 2.0 a0 this term (for ` = 1 in this figure) has
little effect on V (r), so here V1eff (r) V (r). Going the other way, as r decreases to 0 the function V (r) keeps
getting stronger (more negative). But now the barrier term gets stronger more rapidly. By about r 1 a0 , the
barrier term starts to win, and V1eff (r) goes positive. The smaller r is, the more the barrier term dominates the
physical potential energy. By r = 0.01 a0 , the physical potential energy equals 294.623 E h , while the barrier
term is 104 E h , giving an effective potential energy of 9.7 103 E h . Near the origin, the barrier term always
wins.
The right panel shows this battle royal played out for three barriers of increasing strength. As ` increases,
something important happens. Although V1eff (r) remains weakly attractive for r . 1 a0 , the barrier terms for
` > 1 are so strong that V`eff (r) for ` > 1 are purely repulsive. So for this physical potential energy there are no
bound-state solutions to radial equations for ` > 1.

4.5.2 The effective potential energy and the number of bound states
Figure 4.1 shows the great influence of V`eff (r) on the reduced radial functions. So great is this influence that
we can use knowledge of V`eff (r) to answer three crucial questions:
(1) How many bound states of a particular ` does the physical potential energy V (r) support?
(2) For each such bound state, what is the qualitative behavior of un,` (r) for all r?
(3) Where are the classical turning pointsthose special values of r that separate each classically
allowed region from the adjacent classically forbidden region?
Well discuss the qualitative behavior of the radial function in Secs. 4.6 and 4.7. Here Ill use Fig. 4.1 to
discuss the number of bound states (see also Complement 4.C).
According to the reduced radial equation Eq. (4.4.4), for each ` = 0, 1, 2, . . ., there may exist one or
more bound-states. Each bound-state has a discrete energy En,` , and a function un,` (r) that satisfies the
boundary conditions un,` (r) 0 as r 0 and as r . What determines the number of bound states
for a given ` is the effective potential energy V`eff (r). As illustrated in Fig. 4.1, no matter how strong the

JQPMaster

Version: 8.35

Printed: August 17, 2010

288

4.6. Generic properties of the reduced radial function

physical potential energy, there (almost) always exists some `max such that such that V (r) supports no
bound states with ` > `max .26
Key Points
The barrier term results in an effective potential energy V`eff (r) that, for ` > 0, is weaker (less attractive)
than the physical potential energy V (r) at all r.
The qualitative nature of un,` (r) (and hence of the radial probability density) is determined by the competition between the barrier term and V (r) and can depend strongly on the orbital angular momentum
quantum number `.
The physical potential energy determines the total number of bound states of the system. But the
effective potential determines the number of bound states for each orbital angular momentum `.
Since V`eff (r) determines the number of bound states of orbital angular momentum ` for a system,
the number of bound states decreases with increasing `. Typically, there is some `max such that V (r)
supports no bound states for ` > `max the notable exception being the pure Coulomb potential energy
of Chap. 5.

4.6. Generic properties of the reduced radial function


There is a thread here which we have not yet grasped
and which might lead us through the tangle.
Sherlock Holmes in The Adventure of the Devils Foot by Sir Arthur Conan Doyle

Nature is staggeringly diverse. Much of what makes it tractable is that we need not start anew with
every system, because systems share generic properties. For single-particle systems, for instance, every
stationary-state wave function (r, t) has the form
(r, t) = E (r) ei Et/~ ,

b E (r) = EE (r).
where H

(4.6.1)

In the study of rotationally invariant 3D systems, every stationary angular momentum eigenstate (SAME)
has the form
1
E,`,m` (r, , ) = RE,` (r)Y`,m` (, ) = uE,` (r)Y`,m` (, )
(4.6.2)
r
where Y`,m` (, ) is a spherical harmonica function thats known, system-independent, and hence generic.
Moreover, every reduced radial function satisfies the reduced radial equation,
~2 d2
~2 `(` + 1)

u
(r)
+
un,` (r)
n,`
2m dr2
2mr2
|
{z
} |
{z
}
radial KE

angular KE

V (r) un,` (r)


|
{z
}

physical potential energy

En,` un,` (r) = 0


|
{z
}

(4.6.3)

energy

Not only equations, such as the TISE and the radial equation, and functional forms, such as the time
factor ei Et/~ and spherical harmonics, are generic. So are boundary conditions: every bound-state SAME
radial function Rn,` (r) must be constant at r = 0 and go to zero as r .
Because generic properties must hold for large classes of similar physical systems, knowing them greatly
empowers a physicist. Generic properties
(1) simplify the algebraic and/or numeric work required to calculate properties of a system;
(2) guide solution of specific equations, such as the reduced radial equation;
26 Details: An exception to this statement is the pure-Coulomb potential energy of Eq. (4.2.5), p. 269, a very special case for
which, as well see in Chap. 5, an infinite number of bound stationary states exist for all `.

JQPMaster

Version: 8.35

Printed: August 17, 2010

4.6.1 Limiting behavior of physically reasonable potential energies

289

(3) enable (almost) fool-proof checks on results of numerical solutions of equations;27 and
(4) develop insight and the ability to deduce, with little or no algebra, qualitative information about
properties of a system.
The tactic Im advocating combines two of our habits for effective problem-solving (Appendix Q): Begin by
brainstorming and Pursue familiar patterns:

Tip. Before trying to solve an equation, either on paper or on a computer, list every property you know its solutions
must satisfy by virtue of the nature of the system the equation describes. Then use these generic properties to
simplify the equations you must solve and the calculations you must perform, and to deduce as much as possible
about the results you expect. After youve solved the equation, carefully compare your answers to the generic
properties in your list and resolve any discrepancies.

For radial functions, some generic properties follow from requirements such as orthogonality and boundary
conditions. Others follow from the mathematical structure of the reduced radial equation, Eq. (4.6.3).
Because this equation contains V (r), its solutionsthe functions un,` (r) and energies En,` are systemdependent, not generic. To find generic properties, we must look to regions of r where the potential-energy
term V (r)un,` (r) is negligible compared to other terms in this equation. There are two such regions: very
small r and very large r.

4.6.1 Limiting behavior of physically reasonable potential energies


Potential energies in nature are finite and continuous. A typical attractive (negative) spherically symmetric
potential energy has the qualitative properties illustrated in Fig. 4.1, p. 267:28
In the asymptotic limit r , the interaction that gives rise to the potential energy vanishes, so
V (r) approaches zero:29
V (r) 0,
r

asymptotic limit of a physically reasonable potential energy.

(4.6.4)

In the near-zero limit r 0, the potential energy goes to some negative value. (Some model
potential energies, such as the pure Coulomb potential energy well study in Chap. 5, go to in this
limit.)
To illustrate these behaviors, Ive plotted model potential energies for lithium, helium, and the molecule C60
(fullerene) in Fig. 4.1. All these potential energies go to zero (at quite different rates) as r and to some
large negative value (or to ) as r 0.
27 Commentary: For instance, if your computer generates a reduced radial function that violates one of its generic properties,
such as the asymptotic boundary condition un,` (r) 0 or some of the algebraic conditions well discuss in this section, then
you know your answer is wrong.
28 Memory Jog: We choose the zero of energy at the top of the potential well. With this definition an attractive potential
energy is one thats negative everywhere. With this choice, all bound states have negative energies, En,` < 0, and the continuum
consists of all positive energies E > 0. Any potential energy that supports one or more bound states must be attractive for
some range of r. Most potential energies in nature are attractive for all r, with the possible exception of a small region near
r = 0 and/or at very large r.
29 Commentary: Although the interaction V (r) between a particle and a force center may be of infinite range, it must diminish
as the separation between the particle and the force center increases (as r ). For instance, the pure Coulomb potential
energy, which varies with r as 1/r, never equals zero. But as r , the Coulomb potential energy goes to zero: V (r) 0.
Note that some widely-used models display unphysical behavior at large r. For instance, the potential energy of an isotropic
3D SHO, V (r) = m 2 r2 /2 blows up as r . Such potential energies can be used to model interactions for small r only,
because their large-r limiting behavior is unphysical.

JQPMaster

Version: 8.35

Printed: August 17, 2010

290

4.6.2 Cleaning up the reduced radial equation


0.

-r 4
potential energy

Figure 4.1. Large r behavior of some typical physical potential energies. The potential energies shown (in atomic units) are
a model potential for the fullerene C60 (solid
curve), the long-range behavior of the electronatom interaction potential energy for a He atom
in the 2 1 S excited state with polarizability
= 803 a30 (dash curve), and the model Li
potential energy developed in Complement 4.D
(short dash curve). Notice that the horizontal
axis starts at r = 9.0 a0 .

-0.05

Li
-0.1

-0.15

-0.2

C60

10

11

12

13

14

15

4.6.2 Cleaning up the reduced radial equation


Before analyzing the radial equation (4.6.3) at small and large r, I want to eliminate some clutter. First, Ill
temporarily omit the usual subscripts from u(r) and E. Second, Ill multiply everything by 2m/~2 , thereby
isolating the second derivative, which will play a primary role in the analysis. Heres whats left:

2m
d2
`(` + 1)
2mE
u(r) + 2 V (r) u(r) +
u(r) +
u(r) = 0.
2
2
dr2
~
r
| {z }
| ~{z }

(4.6.5a)

U (r)

With the indicated definitions,


U (r)

2m
V (r),
~2

and

2mE
,
~2

(4.6.5b)

this equation simplifies to the clean, convenient form30

d2
`(` + 1)
u(r) + U (r)u(r) +
u(r) + 2 u(r) = 0
dr2
r2

(4.6.6)

where is called the decay constant. I can now generate an asymptotic equationthe reduced radial
equation in the asymptotic limit r .

I Warning: Dont forget that the decay constant depends on n and `that is, each radial function has a
different decay constant. Restoring the subscripts, the definition of the decay constant reads31
r
~2 2n,`
2m(En,` )
n,`
=
E
=

n,`
~2
2m
r
2 2
~ n,`
2mn,`
= n,` =
n,`
2
~
2m

(4.6.7a)
(4.6.7b)

30 Details: Its always a good idea, when manipulating equations, to check that the dimensions are treated correctly. The

dimensions of potential energy, of course, are [V (r)] = E . Since 2m/~2 = E1 L2 , the dimensions of U (r) are [U (r)] = L2 .
This is consistent with the other terms in Eq. (4.6.6). Note also that [] = L1 .
p
31 Jargon: The decay constant is related to the more familiar wave number k
2mEn,` /~2 by kn,` = i n,` . The wave
n,`
number is inconvenient for discussing bound states because its imaginary. (Bound-state energies En,` are negative.) But for a
p
continuum state with energy E > 0, the wave number kn,` 2mE/~2 is real. So scattering theories are usually formulated
in terms of kn,` . The main advantage of the binding energy is that its positive: this feature makes it easier to keep track of
the energy ordering of bound states and eliminates the need to keep track of pesky minus signs.

JQPMaster

Version: 8.35

Printed: August 17, 2010

4.6.3 Generic behavior of reduced radial functions in the asymptotic limit

291

The expression in terms of the binding energy E [Eq. (4.2.3b), p. 269] emphasizes that, since E < 0
for all bound states, the decay constant is real.

4.6.3 Generic behavior of reduced radial functions in the asymptotic limit


4.6.3.1. Derivation of the asymptotic reduced-radial equation
As r , the physical potential energy V (r) goes to zero according toV (r) 0 as r [Eq. (4.6.4),
p. 289]. Similarly, the barrier term ~2 `(` + 1)/2mr2 goes to zero. So for any V (r) is, there is some r large
enough that we can approximate the radial equation (4.6.6), to any desired accuracy, by

d2
u(r) + 2 u(r) = 0,
dr2

asymptotic reduced
radial equation.

(4.6.8)

To see this, look at Fig. 4.2. The solid curve shows the terms retained in the approximation Eq. (4.6.8);
the dashed curve shows all four terms in complete reduced radial equation (4.6.6). As r increases, terms other
than those in Eq. (4.6.8) cease to be important: to graphical accuracy, the two curves are indistinguishable
beyond about 14 a0 .

0.000

-0.002

Figure 4.2. Asymptotic behavior of the the


` = 1 reduced radial equation for boron.
This figure shows large-r behavior of terms in
the reduced radial equation for the 2p electron
in the ground state of boron. The solid curve is
the sum of the second-derivative d2 u2p (r)/ dr2
and 2p u2p (r); the dashed curve is the sum of all
terms in the 2p reduced radial equation. [Calculations based on the model-based valence-electron
(2p) radial functions for boron of Morse et al.
(1935).]

-0.004

-0.006

-0.008

BH2pL
-0.010

10

12

14

16

r Ha0 L

4.6.3.2. Solution of the asymptotic reduced-radial equation


What kind of a function solves Eq. (4.6.8)? It must be a function whose second derivative equals 2 times
the function itself. One of the derivatives you learn in elementary calculus does the job:32
d r
d2 r
e
= er =
e
= 2 er .
dr
dr2

(4.6.9)

So the asymptotic equation (4.6.8) is solved by functions whose radial dependence is exponential, either er
or er . Since Eq. (4.6.8) is a second-order differential equation, its most general solution is an arbitrary
linear combination of these two linearly independent functions:33
32 Details: Any linear combination of e+r and er yields a new set of (two) linearly independent solutions of the asymptotic
radial equation. The most common such set consist of the hyperbolic sine and cosine function, sinh z = (ez ez )/2
and cosh z = (ez + ez )/2. These functions are less convenient than (real) exponential functions, because they less easily
accommodate the boundary conditions for bound states.
33 Notation: I use the symbol to indicate the nature of the variation of a function (or an equation) on its variables in some
limithere, the asymptotic limit r . This notation is different from , which I use to indicate the limiting value of a
quantity, such as u(r) 0 as r .

JQPMaster

Version: 8.35

Printed: August 17, 2010

292

4.6.3 Generic behavior of reduced radial functions in the asymptotic limit


u(r) A er + B er

as r ,

(4.6.10)

with as-yet-undetermined constants A and B. But wait. Asymptotically, all bound-state reduced radial
functions must go to zero. Well, er doesnt cooperate at all: in the limit r , it goes to . This generic
asymptotic boundary condition demands that we set B = 0. Hence the generic asymptotic r-dependence
of a reduced radial function must be
u(r) er 0 as r

asymptotic r dependence of
all reduced radial functions

(4.6.11)

Rule: All bound-state reduced radial functions decay to zero exponentially as r . The rate of decay is
determined by the decay constant defined in Eq. (5.4.3c), which depends on the quantum numbers n and ` of
the state.34
We can develop a similar generic functional form for the other limit where V (r) is negligible: near the origin.
Thats where well go in 4.6.4.

Example 4.8 (Asymptotic decay of reduced radial functions for scandium.)


Scandium (Sc: Z = 21) is a fairly soft, silvery metal. You can find more scandium on a star than on earth, but
on earth youre likely to find it in labs, since physicists find its properties fascinating. Neutral scandium has
21 electrons: 20 core electrons and one valence electron (the Aside on p. 341). The valence electron has
principal quantum number is n = 4 and orbital quantum number ` = 0.
Figure 4.3 shows the limiting behavior of radial wave functions for Sc with principal quantum number n = 3that is, u3s (r), u3p (r), and u3d (r). I calculated the decay constants for these functions from
the corresponding binding energies:35
3,0 = 55.0 eV = 3,0 = 2.011 a1
0

(4.6.12a)

3,1 = 33.0 eV = 3,1 = 0.588 a1


0

(4.6.12b)

3,2 = 8.0 eV = 3,2 = 0.767 a1


0 ,

(4.6.12c)

where for convenience Ive converted the decay constants to units of inverse Bohr (atomic units: Appendix F).
These values show that in Sc the 3s and 3p electrons are much more tightly bound than the 3d electron. Look
carefully at Eqs. (4.6.12) together with the curves in Fig. 4.3. What trends do you spot? How does the rate of
asymptotic decay of uE,` (r) depend on the binding energy and on the decay constant?

Key point! The decay constant for a SAME n,`,m` (r) is a quantitative measure of how rapidly Rn,` (r)
and un,` (r) decay to zero as r . The larger the binding energy n,` , the larger the decay constant n,` , and
the more rapidly these functions decay to zero in this limit.
34 A cautionary note: The asymptotic behavior of a wave functionits behavior as r depends on the asymptotic form
of the differential equation it satisfies. This equation, in turn, depends on the asymptotic behavior of the potential energy in the
Hamiltonian. The generic behavior in Eq. (5.4.5b) does not pertain to a potential energy V (r) that increases without limit as
r . No physical potential energy really behaves this way, but many widely used model potential energies do. An example
is the isotropic 3D harmonic oscillator with natural frequency : that is, V (r) = m 2 r2 /2. The un,` (r) function for this
2

potential energy does decay to zero as r , but it does so like a Gaussian functionas er for = m/2~rather than
as e r . The same caveat pertains to the other properties developed in this section. In the near-zero limit r 0, the generic
behavior of un,` (r) pertains to any potential energy that goes to zero as r 0 like V (r) C/r where < 2. But no matter
what V (r) is, un,` (r) must have nr = n ` 1 nodes in the open interval (0, ).
35 Read on: See Table 1.4 in Atoms and Their Spectroscopic Properties by V. P. Shevelko (New York: Springer, 1997).

JQPMaster

Version: 8.35

Printed: August 17, 2010

4.6.4 Generic behavior of reduced radial functions in the near-zero limit

{=0

293

{=2

0.77

fn,{HrL
{ =1

{ =1

{=2

0r

{=0

1.56
2.01

intermediate r

Figure 4.3. Limiting behavior of the bound-state reduced radial functions for scandium. We consider un,` (r) for ` = 0 (s states), ` = 1 (p states), and ` = 2 (d states). The
values of n,` used in the asymptotic limiting functions Eq. (4.6.8), shown as small numbers
above each curve on the right, were calculated from the binding energies of the n = 3 electrons
in scandium; see Eq. (4.6.12).

4.6.4 Generic behavior of reduced radial functions in the near-zero limit


What happens to the terms in the reduced radial equation (4.6.6) near the origin, as r 0? For one thing,
the barrier term, ~2 `(` + 1)/2mr2 (for states with ` > 0) goes to +. This behavior competes with that of
the physical potential V (r), which is negative and, in general, approaches a large negative number (or ).
Well consider the (very) large class of physical potential energies that, in this limit, can be written as a
series expansion about r = 0 (see the Aside on p. 295)
V (r) =

c1
+ c0 + c1 r + c2 r2 + ,
r

(4.6.13)

where for an attractive potential energy, the coefficients cn are negative or zero. Examples include the pureCoulomb model of a one-electron atom, Eq. (4.2.5), p. 269 (Chap. 5); the lithium model of Complement 4.D;
and the models of Fig. 4.3, p. 279. In the competition between an attractive V (r) and the repulsive barrier
term, as r 0 the barrier term wins. You can see the small-r behavior of all four terms in the reduced
radial equation in Fig. 4.4. As r 0, the only survivors are the second-derivative d2 u/dr2 and the barrier
term [~2 `(` + 1)/2mr2 ]u(r).
4

Figure 4.4. Near-origin behavior of the reduced radial equation for boron. This figure shows the small-r behavior of terms in the
reduced radial equation (4.6.6) for the 2p electron in the ground state of boron. The terms are
identified by labels on the figure. Note that the
energy-dependent term 2 u(r) (the dotted curve)
is so small as to be barely visible on this figure.
These calculations are based on the model-based
valence-electron (2p) radial functions for boron
of Morse et al. (1935).

{H{+1Lr 2 uHrL
2 uHrL

-2 uHrLr 2
-2

-4

UHrL uHrL
-6

BH2pL
-8
0.0

0.1

0.2

0.3

0.4

r Ha0 L

For small enough r, therefore, we can approximate the reduced radial equation Eq. (4.6.6) by

JQPMaster

d2
`(` + 1)
u(r) +
u(r) = 0,
dr2
r2

reduced radial equation


in the r 0 limit (` > 0).
Version: 8.35

(4.6.14)
Printed: August 17, 2010

294

4.6.4 Generic behavior of reduced radial functions in the near-zero limit

This second-order differential equation has two linearly independent solutions: one is proportional to r`+1 ,
the other to r` . The most general solution is therefore
u(r) C r`+1 + D r`

as r 0

(` > 0)

(4.6.15a)

for arbitrary complex constants C and D.


As in our study of the asymptotic form of u(r), we now ask whether either term in (4.6.15a) must be
rejected on physical grounds. Again, one term must go: in this limit, r` blows up! Were left with
near-origin r dependence of
reduced radial functions (` > 0).

u(r) r`+1 0 as r 0,

(4.6.15b)

Im sure you noticed that so far, my argument pertains only to states with ` > 0. For ` = 0, there is no
barrier term, so the reduced radial equation (at any r) is simply

d2
u(r) + U (r)u(r) + 2 u(r) = 0
dr2

reduced radial equation for ` = 0

(4.6.16)

In this case the only physically acceptable (normalizable) solutions are those that, as r 0, go to zero like
u(r) r 0 (see footnote 36). We conclude for all ` that the physically admissible solutions are those
that obey
Rule: As r 0, all bound-state reduced radial functions go to zero like r`+1 .
un,` (r) r`+1 0

as r 0

near-origin r dependence of
all reduced radial functions

(4.6.17)

Unlike the generic asymptotic behavior un,` (r) er 0 as r , this near-zero behavior does not
depend on the energy of the state.
I

Try This!

Example 4.9 (The short-range behavior of radial functions for scandium.)


Look at the near-zero limiting behavior of the n = 3 functions for Sc in Fig. 4.3, p. 293. Only the ` = 0 function
(an s state) approaches zero linearly; the ` = 1 function (p state) goes to zero quadratically, and the ` = 2
rad
function (d state) goes like r3 . Now, u2n,` (r) is a radial probability density Pn,`
(r). So except for s states, the
probability of finding the electron very near the origin is vanishingly small. Moreover, the size of this small-r
region of very low probability grows with increasing `.
J
4.10. Verifying the near-zero solution.
Verify by direct substitution that r`+1 and r` satisfy Eq. (4.6.14). If you want more insight into this
solution, check out the next Aside.

Aside. Details of the derivation of the near-origin behavior the reduced radial function.
Here well take a closer look at the details leading to Eq. (5.4.5a). The function un,` (r) satisfies the reduced
radial equation [Eq. (4.6.6), p. 290]. Being of second order, this equation has two linearly independent
solutions. If, as r 0, the physical potential energy V (r) goes to zero like r for some 0 < < 2, then
in this limit, V`eff (r) (for ` > 0) is dominated by the barrier term. So for any ` > 0 the behavior of the
solution of Eq. (4.6.6) for small enough r wont depend on V (r). In fact, if r is sufficiently small, then
~2 `(` + 1)/2mr2 |En,` |, and the behavior of uE,` (r) wont depend on the energy either! The remaining
small-r equation Eq. (4.6.14) shows how the two linearly independent solutions un,` (r) of the exact equation (4.6.6) must behave as r 0: one must depend on r like uE,` (r) r`+1 , the other like uE,` (r) r` .
As discussed above, we must jettison solutions that blow up and so are left with solutions that obey the
regularity condition uE,` (0) = 0 by going to zero.36

36 Details: For ` = 0, the right-hand side of Eq. (4.6.14) is zero, so u


n,0 (r) goes to a constant as r 0. At first glance,
this behavior seems okay, since it gives a radial function Rn,` (r) that near the origin behaves like RE,0 (r) 1/r. While
this behavior doesnt produce an infinite normalization integral, it does cause trouble in the Schr
odinger equation, because as

JQPMaster

Version: 8.35

Printed: August 17, 2010

4.6.5 Generic behavior between the near-zero and asymptotic limits

295

Aside. The near-zero limit and power-series expansion of the reduced radial function.
The reduced radial equation has a useful mathematical property: the origin r = 0 is a regular singular
point.37 This property implies that Eq. (4.6.14) admits a power-series solution. So if necessary we can
obtain a more detailed mathematical description of the smallr behavior of its solution via the expansion38

X
s
i
(4.6.18)
un,` (r) r
c0 +
ci r , with c0 6= 0.
r0

i=1

4.6.5 Generic behavior between the near-zero and asymptotic limits


Weve discovered that the reduced radial function for every bound state of a rotationally invariant system
must obey the limiting behaviors39
un,` (r) r`+1 ,

and un,` (r) en,` r .


r

r0

(4.6.19)

Between the extremes of Eq. (4.6.19)for intermediate values of rthe function un,` (r) depends on the
potential energy. So you might think we cant deduce anything generic about this region. But we can.
To describe the functional dependence of un,` (r) in this intermediate region, Ill define an auxiliary
function fn,` (r) so as to incorporate the limiting behaviors of Eqs. (4.6.19):
un,` (r) = Nn,` r`+1 fn,` (r) en,` r

(4.6.20)

where Nn,` is a normalization constant. We could derive an equation for fn,` (r) (see Exercise 4.10) which we
could solve by expanding fn,` (r) in a power series in r. Instead, lets see what generic properties of fn,` (r)
we can discover.
Generic properties of the intermediate-r function. Imagine that we were to expand fn,` (r) in a
power-series in r:

X
X
fn,` (r) =
aj rj = a0 +
aj rj .
(4.6.21a)
j=0

j=1
`+1

n,` r

The product of this series times the limiting forms r


and e
must not violate the boundary conditions
on un,` (r). For example, in the r limit, fn,` (r) must not go to faster than en,` r goes to zero. To
guarantee this, the number of terms in Eq. (4.6.21a) must be finite:

as r ,

jX
max
j=0

aj r j

{z

wants to

er

{z

0,

if jmax is finite.

(4.6.21b)

wants to 0

r 0, the kinetic-energy operator acting on a function that behaves like RE,0 (r) 1/r produces a Dirac delta function, 3 (r)
[see Appendix N]behavior that is even more unacceptable than an infinite normalization integral! [This function appears
because 2 (1/r) = 4 3 (r).]
37 Jargon: A second-order linear differential equation of the form
d2 f (r)
df (r)
+ g(r)
+ h(r)g(r) = 0
dr2
dr
is regular at r = 0 if both g(r) and h(r)
differentiable) at r = 0. For the reduced radial equation, r = 0
are analytic (infinitely

is a regular singular point. because limr0 r2 h(r) < .


38 Commentary: This strategy works whether or not we can derive a recursion relation for the expansion coefficients suitable
for all r. As r 0 the factor rs dominates un,` (r), so our conclusions hold even if we must solve the radial equation numerically
(see Chap. 5 for an exception).
39 Commentary: The class of spherically symmetric potential-energy functions to which these behaviors pertain is very large;
but physicists sometimes use model potential energies whose properties exclude them from this class. A prominent example is the
isotropic three-dimensional spherical harmonic oscillator, V (r) = m02 r2 /2. Unlike any potential energy in nature, this function
goes to infinity as r behavior that alters the asymptotic form of its radial functions. To find out about restrictions on
JQPMaster
8.35
Printed: August 17, 2010
this class of functions, see Chap. 6 of Galindo and PascualVersion:
(1991).

296

4.6.5 Generic behavior between the near-zero and asymptotic limits

Similarly, in the near-zero limit r 0, the auxiliary function fn,` (r) must not cause un,` (r) to deviate from
the required r`+1 behavior. To guarantee this, we require that the constant term a0 in Eq. (4.6.21a) be
non-zero. Such conclusions, which required thought rather than algebra, are a boon in actual solution of the
radial equation, as youll see in Chap. 5.
Much ado about node things. We can deduce one further generic property of fn,` (r) and hence
of un,` (r). As discussed in Secs. 4.2.64.2.7 and illustrated in Example 4.2, orthogonality of the functions un,` (r) with different n and the same ` requires that each un,` (r) has nr = n ` 1 nodes positioned
so hun0 ,` |un,` ir = 0 for n0 6= n for all functions unr ,` (r) with that ` [Eq. (4.2.14b), p. 272]. Since neither r`+1
nor en,` r has nodes, those n ` 1 nodes must be a property of fn,` (r):
number of nodes in fn,` = n ` 1.

nodal structure of the


auxiliary radial function.

(4.6.22)

Key Points
Bound-state reduced radial functions decay exponentially to zero in the asymptotic limit as en,` r .
The rate of decay of un,` is controlled by the decay constant n,` , which depends on the bound-state
energy, since En,` = n,` = ~2 2n,` /2m. The more tightly bound the electron, the more rapidly un,`
decays to zero as r .
Bound-state un,` (r) decay to zero as r 0 like r`+1 . This behavior doesnt depend on the energy.
The reduced radial function for any bound state must have the form
un,` (r) = r`+1 (finite polynomial in r) en,` r

(4.6.23)

where the polynomial has n ` 1 nodes in 0 < r < . If your computer spits out a radial function
for a spherically symmetric potential energy that does not satisfy this form, then you (who are smarter
than your computer) know its made a mistake.

JQPMaster

Aside. On the use of mathematical expedients in quantum mechanics.


Reduced radial functions and effective potential energies are so much a part of day-to-day quantum mechanics that its easy to loose sight of their peculiar status. The reduced radial function un,` (r), in particular,
is nothing more than a mathematical artifice we introduced to simplify determination of the radial function Rn,` (r), which appears in the SAME Rn,` (r)Y`,m` (, ). The effective potential energy V`eff (r) contains
the barrier term ~2 `(` + 1)/2mr2 , which appears in the differential equations for both Rn,` (r) and un,` (r).
Although this term is part of V`eff (r), which we call a potential energy, its nothing of the sort: the barrier
term arises from the angular kinetic energy operator (4.5). Yet, because of its role in the reduced radial
equation, the effective potential is largely responsible (along with the boundary conditions and orthogonality
requirements) for the qualitative features of the functions { un,` (r) } for a system with physical potential V (r).
rad
And, because the radial probability density Pn,`
(r) = u2n,` (r), the function V`eff (r) greatly influences the radial distribution of position probability densitywhich in turn influences almost every physical property of
the particle.

Version: 8.35

Printed: August 17, 2010

297

4.7. Qualitative solution of the radial equation


The sound of physics.
The soft, breathless whir of Now.
Just listen.
Close your eyes, pay attention. . . .
A purring electron? Photons, protons?
Yes, and the steady hum of a balanced equation.
The Nuclear Age by Tim OBrien

In this section, well use our insights into the generic properties of reduced radial functions (4.6) to solve
their equation qualitatively.

Tip. Always generate a qualitative solution of the radial equation before trying to solve it quantitatively.

Youll often find that a qualitative solution contains the information you need, and you neednt bother with
a quantitative solution.40
Essential plots. We start with a graph of the physical and effective potential energies like the left panel
in Fig. 4.1, p. 287. Well need one such graph for each ` for which we think V (r) might support one or
more bound states (see 4.5.2). These days, its reasonably easy to efficiently generate such graphs on
a modest computer.41
Essential energy estimates. We next estimate where the energies En,` for SAMEs with each ` lie on the
graph of V`eff (r). We do this using The Principle of Pattern Seeking (Appendix Q). Heres how. We
compare our V (r) to the potential energy of a system for which we know (or can look up) the bound-state
energies. For an atom, we could start with the bound-state energies of an electron in a one-electron atom
with nuclear charge +Ze (Chap. 5):42
En =

Z2
E h,
2n2

(for a one-electron atom.)

(4.7.1)

where E h = 27.212 eV. Lets see how to apply this gambit to atomic lithium.
I

Example 4.10 (Estimating the energy of a 2s electron in the ground state of lithium.)
In Complement 4.D I discuss the following model potential energies for the valence (2s) electron in the groundstate of lithium (Li: Z = 3), [Eq. (4.D.4b), p. 341],
V2s (r) =

1
( r + 2) r

e
,
r
r

model potential energy for


the 2s electron in Li,

(4.7.2)

where Z = 3 and = 5.38 a1


0 . The valence electron has n = 2 and ` = 0its a 2s electron. (You dont need
to have read Complement 4.D to understand this example). The left panel in Fig. 4.1 compares Eq. (4.7.2) to
pure-Coulomb potential energies [Eq. (4.2.5), p. 269] for Z = 1 (a hydrogen atom) and Z = 3 (a one-electron
Li++ ion). For each of these comparison potentials, the energies are given by Eq. (4.7.1). To estimate the
energy E2s for Li, Ill start with the energy of an n = 2 state of atomic hydrogen, then take into account
alterations in the H-atom potential energy that make it more like Eq. (4.7.2). I prefer positive quantities, so
Ill use the binding energy 2s E2s .
40 Commentary: You can also apply the techniques of this section [except for applying the asymptotic boundary condition
un,` (r) 0] to continuum states (see Complement 4.B).
41 Read on: Several shareware packages you can download off the internet will quickly and easily generate attractive graphs
of either numerical or analytical functions. One of the most widely used is GNUplot. If, after trying out a shareware package,
you decide to keep using it, dont forget to pay the (very modest) registration fees! Its the least you can do for the people who
worked so hard to develop the software. Of course, more high-powered software like Excel, Mathematica, and graphics-specific
packages can also do the job.
42 Read on: This is the bare-nucleus model of 4.D.3 in Complement 4.D; see also Chap. 11.

JQPMaster

Version: 8.35

Printed: August 17, 2010

298

4.7.1 Classically allowed and classically forbidden regions

30.610 eV

Li++

Energy HeVL

-10

potential energy

-1r
Li
-20

-3r

-30

-40

-50
0.0

0.1

0.2

0.3

0.4

0.5

5.400 eV

Li

3.401 eV

0.6

Figure 4.1. Comparison of a model Li potential energy to pure-Coulomb


potential energies. The solid curve shows the model Eq. (4.7.2). The other curves
show pure-Coulomb potential energies [Eq. (4.2.5), p. 269] for a one-electron atom with
Z = 1 (atomic hydrogen) and with Z = 3 (doubly ionized lithium). The right panel
compares my estimates of the 2s binding energy for lithium against the numerically
calculated value for Eq. (4.7.2). Energies are in eV. See also Complement 4.D.

For Z = 1 and n = 2, Eq. (4.7.1) gives 2s = 0.125 E h = 3.40 eV. According to Fig. 4.1, my model V (r)
for a 2s electron in Li is stronger (more attractive) than the Coulomb potential energy for hydrogen. So 2s
for Li must be larger than 2s for H: that is, 3.40 eV is an lower bound on 2s for Li.
I can also estimate an upper bound on this energy. For the Li++ ion Eq. (4.7.1) with Z = 3 gives a
2s binding energy of 30.6 eV. Fig. 4.1 shows that the pure-Coulomb potential energy for Z = 3 is stronger
than the model V (r) for Li. Therefore 30.6 eV is a upper bound on 2s . I conclude that the true binding energy
lies in the interval 3.4 eV < 30.6 eV < 2s >. Looking closely at Fig. 4.1, I speculate that the true 2s will be
closer to 3.4 eV, than to 30.6 eV. Sure enough, numerical solution of the TISE for my model potential energy
gives 2s = 5.4 eV. The bounds and correct energy are compared in the energy-level diagram in the right panel
of Fig. 4.1.
J

4.7.1 Classically allowed and classically forbidden regions


Armed with a plot of V`eff (r) and a guess at n,` , we can determine the qualitative behaviorthe shape
of un,` (r). Our goal is a qualitatively accurate sketch of this function. First we must classify regions of r
according to the qualitative behavior of un,` (r) in each region.43
To explain what Im talking about, I must (yet again) rewrite the reduced radial equation Eq. (4.4.4),
p. 282, this time as

2
2m eff
2m
d
2 V` (r) + 2 En,` un,` (r) = 0,
(4.7.3a)
dr2
~
~
where V`eff (r) V (r) +

~2 `(` + 1)
.
2mr2

(4.7.3b)

Im now going to write Eq. (4.7.3a) in terms of the local wavenumber


43 Read on: Chap. 9 of Understanding Quantum Physics develops the tools for understanding the qualitative behavior of
a 1D bound-state wave function in detail, with lots of examples.

JQPMaster

Version: 8.35

Printed: August 17, 2010

4.7.1 Classically allowed and classically forbidden regions


r

2m
En,` V`eff (r)
~2

local wavenumber.

(4.7.4)

d2
2
+
k
(r)
un,` (r) = 0,
n,`
dr2

(CA region)

(4.7.5)

kn,` (r)

299

Unlike an actual wave number, kn,` (r) is a function of r, not a constant.] Lets consider Eq. (4.7.5) in the
two regions of r suggested by the definition (4.7.4): where En,` > V`eff (r), and where En,` < V`eff (r).
Classically allowed regions. Suppose r is such that En,` > V`eff (r). Then En,` V`eff (r) is positive,
2
so kn,` (r) is real. For positive kn,`
(r), the solutions of Eq. (4.7.5) are oscillatory functions. Because a
classical particle can be found only in regions where its kinetic energy is non-negative, its total energy must
exceed or equal its potential energy; thats why we call a range of r values where En,` > V`eff (r) a classically
allowed region (CA).44
Classically forbidden regions. If r is such that En,` < V`eff (r), then En,` V`eff (r) is negative, so kn,` (r)
is imaginary. In this case, its useful to define the local decay constant
kn,` (r) = i n,` (r)

(4.7.6a)

r
= n,` (r) =

2m
2 V`eff (r) En,` = i
~

d2
2
n,` (r) un,` (r) = 0
dr2

2m eff
V` (r) En,`
2
~

(CF region)

(4.7.6b)

(4.7.6c)

Since a classical particle can never enter such a region (because its kinetic energy would be negative),
we call a range of values of r where En,` < V`eff (r) a classically forbidden region (CF). The solutions
of Eq. (4.7.6c) for positive 2n,` (r) are evanescent (decaying) functions of r.45 Remarkably, the key qualitative
difference between the physics of CA and CF regions arise from a mere difference in sign: Eq. (4.7.5) contains
2
+kn,`
(r), with a plus sign, and Eq. (4.7.6c) contains 2n,` (r), with a minus sign.

I Warning: The function that determines whether a region is classically allowed or classically forbidden is
the effective potential, not the physical potential.
Figure 4.2 illustrates these definitions and the behaviors of un,` (r) for the 2s radial function of lithium
(Complement 4.D).
44 Commentary: In the simplest case, an ` = 0 state of the 3D square-well potential V (r) = V for 0 r r with V > 0,
0
0
0
p
the quantity kn,` (r) is the actual wave number kn,` = En,` , and the solutions of Eq. (4.7.5) are sine and cosine functions
of (kn,` r).
45 Jargon: Strictly speaking, an evanescent function decays with increasing distance from the classical turning point (4.7.2).
I use evanescent rather than decaying because to many students decaying implies exponentially decay, that is ei n,` r .
But un,` (r) decays exponentially only when n,` is constant as, for example, in the 3D square well.

JQPMaster

Version: 8.35

Printed: August 17, 2010

4.7.2 Classical turning points

Figure 4.2. Classically allowed and classically


forbidden regions for the 2s reduced radial function in Li. The function u2s (r) (long-dashed curve)
was calculated for the Li model potential energies
(thick solid curve) of in Complement 4.D. The tics
on the vertical axis refer to this function, not V (r).
The horizontal line at 0.198 E h is the energy E2s .
The classical turning point indicated by the vertical
line at r = 5.04 a0 , separates the CA and CF regions.

reduced radial function Ha-12


L
0

300
1.0

u 2 S HrL
0.5
0.0
E2S

-0.5
-1.0

CA

CF

-1.5
-2.0

r2s

10

r Ha0 L

4.7.2 Classical turning points


Imagine following the r axis outward from the origin for some bound-state. As in Fig. 4.2, we pass through
alternating classically allowed and classically forbidden regions. A values of r that separates a CA region
from an adjacent CF region is called a classical turning point.46 The classical turning points are values
of r where the total energy En,` equals the effective potential V`eff (r)that is, theyre solutions of47
En,` = V`eff (r) = V (r) +

~2 `(` + 1)
2mr2

defining equation for classical turning points.

(4.7.7a)

Inserting the definition of V`eff (r) and of n,` , the equation for these points becomes

`(` + 1) + r2 U (r) + 2n,` = 0

where 2n,`

2m
En,` .
~2

(4.7.7b)

This equation is quadratic in r and so has two roots. These roots are called left and right classical turning
ctp
L
R
points and are denoted rn,`
and rn,`
. (To refer to both points at the same time, Ill use the symbol rn,`
.)
For a given physical potential energy V (r) that can support one or more bound states of given `, the
effective potential energy V`eff (r) may have either one or two classical turning points:
R
For an s state (` = 0), the function V0eff (r) defines one nonzero turning point, rn,`
. (The left classical
L
turning point rn,` is at r = 0.)
L
R
For a state with ` > 0, the function V`eff (r) defines two turning points, rn,`
and rn,`
.

Example 4.11 (Classical turning points for a SAME of an atom.)


If we know the analytic form of V (r) and a reasonable guess at the energy En,` , we can calculate the classical
ctp
turning points rn,`
by solving Eq. (4.7.7b). [If, as is often the case, all we have are tabulated values of V (r),
then we can fit these values using, say, a multiparameter regression, or we could solve Eq. (4.7.7b) numerically.]
Heres an example. For a bound electron in a hydrogen atom in the pure-Coulomb model [Eq. (4.2.5),
p. 269] (Chap. 5), the bound-state energies are given by En = Z 2 /2n2 E h , where 1 E h = e02 /a0 = 27.2114 eV.
(Atypically, these energies dont dependent on `.) Setting Z = 1 (atomic hydrogen) and solving Eq. (4.7.7b)

46 Details: If the potential has a bump in it, there may be a CF region that doesnt include either r = 0 or the asymptotic
limit r . If so, then this region is bracketed by classical turning points. As we enter this CF region from the left, the
radial function begins to decay towards the r axis. With further increases in r it will either rise again and pass through the
right turning point into the adjacent CA region or it will cross the axis and enter the CA region on the other side. In either
case, the function behaves like an evanescent function, not like an oscillatory function.
47 Jargon: At a classical turning point a classical particle would have zero kinetic energy; it would turn around and proceed
docilely back into the classically allowed region. The more free-wheeling quantum particle tunnels with impunity into the CF
region. (So dont take the name classical turning point literally.)

JQPMaster

Version: 8.35

Printed: August 17, 2010

4.7.2 Classical turning points

301

for the roots, gives


p

L
rn,`
= n n n2 `(` + 1) ,
(for a SAME of atomic hydrogen with ` > 0).

R
rn,`
= n n + n2 `(` + 1) ,

(4.7.8a)

R
For ` = 0 (an s state), Eq. (4.7.7b) gives only one non-zero solution, rn,0
= 2n2 .
Unfortunately, nature rarely provides systems with a simple analytical potential energy function. Worse,
we almost never know En,` . So a more realistic example is our model Li potential energy Eq. (4.7.2), p. 297].
Inserting this expression into Eq. (4.7.7b) for an s state of Li (Z = 3) gives

E2s

1
1 + er (2 + r) = 0.
r

(4.7.9)

ctp
For this state, the 2s energy is E2s = 0.198 E h . Numerical solution of Eq. (4.7.9) then gives r2s
= 5.04 a0 .
ctp
This turning point is shown as a vertical line on Fig. 4.2. To the left of r2s , the function u2s (r) oscillates,
ctp
ctp
while to the right of r2s
this point, it decays to zero with increasing |r r2s
|. This behavior typifies what an
s-state radial functions does in CA and CF regions.
J

Aside. Amplitude variation in a classically allowed region.


Often a careful sketch of the physical and effective potential energies, a reasonable guess at En,` , and calculated classical turning points will suffice to draw a pretty accurate sketch of un,` (r). Thorough study of oneelectron atoms in Chap. 5 will prepare you to sketch approximate functions for other atoms (see Chap. 11),
because qualitatively the physical potential energy of an atomic electron does not vary appreciably from atom
to atom. If need be, you can generate an even more accurate sketch of un,` (r) by figuring out how its
amplitude, |un,` (r)|, varies with r.
Heres the idea. In a CA region, |un,` (r)| depends on the curvature of the reduced radial function,
the first term d2 un,` (r)/ dr2 in Eq. (4.7.3a); that is, on how rapidly un,` (r) turns over as r changes. A
pointwise measure of the rate of turnover is the local wavelength, which is defined in terms of the local
wave number [Eq. (4.7.4), p. 299] as48
n,` (r)

2
,
kn,` (r)

local wavelength (CA region only).

(4.7.10)

At values of r where the local wavelength n,` (r) is small, un,` (r) turns over very rapidly with increasing r
and so cant get very large before it turns over and heads back towards the horizontal (r) axis. If n,` (r) is
large, however, un,` (r) turns over more gradually, and its amplitude can get larger than for small n,` (r).
Rule: The amplitude of un,` (r) increases as the local wavelength n,` (r) increases. While the maximum
amplitude un,` (r) attains in a CA region depends on its value(s) at the classical turning point(s), this maximum
amplitude is usually larger where n,` (r) is large.
The concept of curvature explains the variation in local wavelength and amplitude of the 2s radial
function in Fig. 4.2, p. 300. Throughout the CA region, 2s (r) is real and positive. Near the origin, 2s (r)
is very small. As r increases to about 1.5 a0 , the function u2s (r) turns over and heads back towards the r
axis. With further increases in r, as u2s (r) crosses the axis, 2s (r) increasesfirst slowly, then very rapidly.
Correspondingly, the rate of turnover of u2s (r) decreases, leading to its long, languid decay to zero as r
in Fig. 4.1.

Key Points
Heres a cookbook for sketching un,` (r) for a particular `:
(1) Sketch V`eff (r), indicating roughly where you think the bound-state energies lie.
(2) Estimate (or, if feasible, calculate) the classical turning points that demarcate the CA and CF regions.
48 Jargon: The magnitude of
n,` is proportional to the magnitude of the curvature of un,` (r). For the ground state, which
has the largest binding energy, u1,0 (r) has the smallest curvature. The radial function for the first excited state has the next
largest smallest curvature, and so on. (For more information, see 9.2 in Morrison (1990) and especially Fig. 9.3).

JQPMaster

Version: 8.35

Printed: August 17, 2010

302

4.8. Final thoughts: tips for potential modelers


(3) Draw a rough sketch of un,` (r) using the guidelines in Tbl. 4.5, p. 306 to treat each CA and
CF region.
(4) Modify your sketch (if necessary) to ensure proper behavior of un,` (r) in the near-zero and asymptotic limits and to ensure that your function has the correct number of nodes.
(5) Use the requirements of continuity of un,` (r) and of its first derivative to draw a smooth curve
through each classical turning point.
(6) If you need a more accurate representation of un,` (r), use the local wavelength, kinetic energy, and
decay constants to adjust the rate of turnover, rate of decay, and amplitude in your sketch.

4.8. Final thoughts: tips for potential modelers


Perhaps it seems surprising that physicists seek beauty.
But in fact they have no choice.
As yet there has not been an exception to the rule that
the demonstrable solution to any problem will turn out to be an esthetic solution.
Gut Symmetries, by Jeanette Winterson (1977)

Two things justify the time weve spent on central-force problems in 2D and 3D. First, the quantum
mechanical analysis of central-force systemsthe progression from symmetry to CSCO to SAMEs to the
reduced radial equationtypifies how physicists approach other (much) more complicated systems. Second,
many systems can be accurately modeled by spherically symmetric potentials (Complement 4.D). In fact,
Chap. 5 is devoted entirely to one such model, the pure-Coulomb model for a one-electron (hydrogenic)
atom, and Chap. 11 generalizes this model to arbitrary atoms.
Here are some suggestions to get you started constructing your own models:
(1) Start with the symmetry properties of the system. Incorporate as many of these properties as possible
into your model. If, for example, the system exhibits axial symmetryas does a diatomic molecule
with respect to the axis defined by its nucleithen you want a model potential energy that doesnt
depend on . Similarly, if the system is symmetric with respect to inversion through some point in
space, then you want a model with this property.
(2) Decide at the outset what physical properties you want to study. You may seek only bound-state
energies. Or you may want Hamiltonian eigenfunctions to use in calculating other properties such as
the average radius. Or you may be interested in a non-stationary state, in which case you need to
construct a wave packet.
(3) Decide how accurately you need these properties. In many real-world applications, youll need only a
rad
good qualitative information, which you can get from sketches of Pn,`
(r) the radial probability density
and comparisons to familiar systems. In other applications, youll need quantitative information, in
which case youll need to solve the reduced radial equation (numerically or analytically) for your model.
(4) Formulate a precise mathematical statement of the problem you want to solve. Choose a coordinate
system that reflects the symmetry properties of the system. Identify any constraints or special processes
that might occurfor instance, will the particle remain bound? Or could the system come apart, as a
molecule dissociates or an atom ionizes?
(5) Most importantly, decide which interactions youre going to include in your model Hamiltonian, keeping
b + V , which necessarily
b=T
in mind that your choices will influence the CSCO. To even write down H
appears in the CSCO, you have to know V . The Hamiltonian is the mathematical articulation of the
story inherent in your model. The system itself is almost certainly quite complicated (otherwise you
wouldnt need a model). So how are you going to treat various complexities of the system? Are the
particles moving very fast, in which case you may need to use special relativity? Can you treat some
of the particles as point particles, perhaps of infinite mass? Do you need to worry about the effects of
spin (Chap. 6)? And so forth.

JQPMaster

Version: 8.35

Printed: August 17, 2010

303
(6) After choosing a Hamiltonian, try to find CSCO. Then figure out the names of the stationary angular
momentum eigenstates. To accomplish this you must address such question as, What are the constants
b make excellent candidates for
of the motion of the system? These operators, which commute with H,
your CSCO. (Keep in mind that not all constants of the motion commute with one another, so not
b
all constants of the motion are eligible for inclusion in a CSCO.) Since the CSCO must include H,
the other operators in it will probably depend on your model. (For instance, if your system is not
rotationally invariant but your model is, then your CSCO can include b
L 2 and b
Lz .) If you later refine
b (If you decide include terms in H
b that
your model, you may need to revise the CSCO as well as H.
arent spherically symmetric, you must jettison operators from your CSCO that commute only with a
spherically symmetric potential energy.)
(7) If you must retain some complicated interactions in your Hamiltonian, consider treating them as perturbations. You can usually identify before doing any calculations that some interactions are too important
to jettison but are clearly weaker than others. If so, plan to incorporate these interactions later using
perturbation theory (Chap. 14 and Chap. 17).
(8) As you devise and then refine your model, dont be rigid. Play around. And remember that your model
need not replicate every feature of the system. For instance, in light of my exhorting you to incorporate
symmetry, you may be wondering, If symmetry is so important, isnt it a bad idea to construct a
model that incorporates symmetries the system does not possess, like modeling a non-spherical system
by a spherical potential? Not necessarily. Your model is supposed to make the Schrodinger equation
easier to solve. If, say, the potential energy isnt really spherically symmetric, but you think the
non-symmetric contributions may be comparatively unimportant for your needs, then by all means
try a spherically symmetric model. (This approximation is widely used in the study of closed-shell
many-electron atoms, which well discuss in Chap. 11.)
(9) Always solve a problem qualitatively before try to solve it quantitatively. Use your experience, intuition, knowledge of general principles, and comparison to other systems youve studied to make
order-of-magnitude estimates of physical properties, sketch eigenfunctions, determine the behavior of
eigenfunctions and physical properties in various limits, and so on. Often dimensional analysis (Appendix B) can yield quite accurate estimates of many physical properties.
Building models of quantum systems is almost as much fun as building models of ships or airplanes. Once
you know a bit of analytic and mathematical machinery, this process can be quite exciting, as you apply
your ever-growing insight to explore nature in the challenging domain of the microverse.

4.9. Users guide to Chap. 4


Its these small details that count. . .
Always the damned details.
The Club Dumas,
by Arturo P
erez-Reverte

Most of the notation used in this chapter appears in the Users guides to Chaps. 13. Table 4.1 lists new
notation and extensions to 3D of notation from earlier chapters.
Symmetry and separation of variables in the central-force problem. Table 4.2 explicitly connects
these symmetry properties and separation of variables for central and non-spherically symmetric potentials, while Tbl. 4.3 relates symmetry to whether the potential commutes with various angular momentum
operators.

JQPMaster

Version: 8.35

Printed: August 17, 2010

304

4.9. Users guide to Chap. 4

nr

radial quantum number ; the number of nodes in Rn,` (r) and un,` (r): nr = n ` 1

n,`,m` (r)

Hamiltonian eigenfunction for a SAME in 3D; in Dirac notation: |n ` m` i

Rn,` (r)

radial function for the SAME n,`,m` (r) = Rn,` (r)Y`,m` (, )

un,` (r)

reduced radial function un,` (r) = r Rn,` (r) = Rn,` (r) = un,` (r)/r

En,`

bound-state energy of a SAME with Hamiltonian eigenfunction n,`,m` (r)

n,`

binding energy n,` En,`

Nn,`
n,`

normalization constant for the radial function Rn,` (r)


p
decay constant n,` 2m|En,` |/~2 ; controls the rate of asymptotic decay of un,` (r)

V`eff (r)

effective potential energy V`eff (r) V (r) + ~2 `(` + 1)2 /2mr2 for physical V (r)

rad
Pn,`
(r)

rad
Pn,`
[ra , rb ]

rad
2
radial probability density for a SAME: Pn,`
(r) = r2 Rn,`
(r) = u2n,` (r)

integrated probability density for a SAME for the interval ra , rb

ctp
rn,`

classical turning point : separates a CA region from an adjacent CF region

hrin,`

mean value of the distance from the origin of a particle in a SAME

(r)n,`

uncertainty in the distance from the origin of a particle in a SAME

Table 4.1. Notation for solution of the central-force problem in D. Also see similar tables in Chaps. 13.
Jargon is set in bold face. In the definition of a classical turning point, CA stands for classically allowed, and CF for
classically forbidden (see Tbl. 4.5).

Symmetry

V dependence

b b
H,
L2

b Lz
H,

separable E (r, , )?

spherical

V (r)

=0

=0

R(r)Y`,m` (, )

axial

V (r, )

6= 0

=0

F (r, )m` ()

none

V (r, , )

6= 0

6= 0

out of lucka

Separability does hold for special forms of V (r).

Table 4.2. Symmetry and separation of variables for single-particle potentials.


potential

symmetry property

commutation relations

V (r, , )
V (r, )
V (r)

none
cylindrical (axial) symmetry
spherical symmetry

none
commutes with b
Lz but not b
Lx or b
Ly

commutes with each component of L

Table 4.3. The relationship between the symmetry properties of the system,
the dependence of the potential energy V (r) on spherical coordinates r, , and
, and the orbital angular momentum operators with which V (r) commutes.

SAMEs and the radial equation for a rotationally invariant system. The key results of this chapter
concerns the radial functions in SAMEs. The radial function satisfies the radial equation
rad

b + V eff (r) Rn,` (r) = En,` Rn,` (r),


T
` = 0, 1, . . . ; and n = 1, 2, . . . ,
(4.9.1a)
`
where V`eff (r) = V (r) +

JQPMaster

~2 `(` + 1)
,
2mr2

Version: 8.35

effective potential for `.

(4.9.1b)

Printed: August 17, 2010

305
The barrier ~2 `(` + 1)/2mr2 in the effective potential energy V`eff (r) in significantly influences Rn,` (r)
for ` > 0. For convenience we actually solve for the reduced radial function
un,` (r) = r Rn,` (r) = Rn,` (r) =

1
un,` (r),
r

(4.9.2a)

a solution of the reduced radial equation

~2 d2
+ V`eff (r)un,` (r) = En,` un,` (r),
2m dr2

` = 0, 1, . . . ;

and

n = 1, 2, . . . .

(4.9.2b)

Bound-state functions un,` (r) must also obey the boundary conditions in Tbl. 4.7, p. 307. Regularity
of un,` (r) at the origin ensures that Rn,` (0) is constant. We can adapt what we know about solving the
TISE in 1D to the reduced radial equation in 3D, provided we keep in mind the differences summarized
in Tbl. 4.4.
One dimension
domain of independent variable

< x <

Three dimensions
0r<
un,` (r) 0

boundary conditions (bound states)

r0

n (x) 0
x

un,` (r) 0
r

the potential energy term

V (x) is the physical potential energy

V`eff (r) is the sum of V (r) and the


barrier term

Table 4.4. distinctions between the time-independent Schr


odinger equation for a 1D potential energy V (x)
and the reduced radial equation for a spherically symmetric V (r) with ` 0.

The parity of a SAME.

Each SAME is an eigenfunction of the parity operator:

b n,`,m (r) = (1)` n,`,m (r)

`
`

b |n, `, m` i = (1)` |n, `, m` i .

(4.9.3)

The parity of a SAME is determined entirely by the parity of its angular function, the spherical harmonic Y`,m` (, ).
The generic form of a bound-state reduced radial function. The function un,` (r) for any bound
state of a potential energy that satisfies the restrictions in 4.2.1 has the generic form
un,` (r) = Nn,` r`+1 fn,` (r) en,` r

generic form of the reduced radial funtion,

where Nn,` is a normalization constant that ensures


Z
|un,` (r)|2 dr = 1,
normalization of the reduced radial function,

(4.9.4a)

(4.9.4b)

and fn,` (r) is a polynomial of degree nr = n ` 1, which therefore has nr nodes.

Rule: To ensure orthogonality of radial functions with different radial quantum numbers nr but the same angular
momentum quantum number `, each function Rnr ,` (r) must have nr nodes, where nr = 0 for the minimum-energy
function, nr = 1 for the function with the next highest energy, and so on [EqPageradialNodesnrad]
number of nodes in Rnr ,` (r) = nr = n ` 1

JQPMaster

Version: 8.35

(4.9.4c)

Printed: August 17, 2010

306

4.9. Users guide to Chap. 4

These nodes must be positioned along the r axis, 0 < r < , so the orthogonality condition Eq. (4.2.14b) will
be satisfied by all radial functions Rnr ,` (r) with that `.
The decay constant n,` controls the
p rate at which un,` (r) decays to zero as r and is related to
the energy En,` of the SAME by n,` 2m|En,` |/~2 . The absolute value bars are necessary because for a
bound state En,` 0. Table 4.5 summarizes the qualitative behavior of un,` (r) in CA and CF regions.
type of region

V`eff (r) versus En,`

CA

V`eff (r) < En,`

CF

V`eff (r) > En,`

differential equation
2

d
2
+
k
(r)
un,` (r) = 0
n,`
dr2
2

d
2

(r)
un,` (r) = 0
n,`
dr2

behavior of un,` (r)


oscillatory
evanescent

Table 4.5. The behavior of the reduced radial function un,` (r) in classically allowed (CA)
and classically forbidden (CF) regions. The local wave number (used in CA regions only) is
defined by Eq. (4.7.4). The local decay constant (CF regions only) is defined by Eq. (4.7.6a).

Angular momentum and the radial equation for SAMEs. Table 4.6 collect various quantities weve
defined in our investigations of the TISE for rotationally invariant systems . Crucial properties of the radial
function for bound states appear in Tbls. 4.7 and 4.8; these are contrasted to the properties for continuum
states in Tbl. 4.9. To aid in solving problems, Ive summarized the properties of the radial function in
classically allowed and forbidden regions in Tbl. 4.10 (see also Tbl. 4.5).
CSCO = SAMEs exist

b b
L 2 , Lz }
{ E,`,m` (r) } are eigenfunctions of { H,

CSCO = can separate radial and angular variables

E,`,m` (r) = RE,` (r)Y`,m` (, )

CSCO = can separate and

Y`,m` (, ) = `,m` ()m` ()

Chap. 2 = solutions of Lz eigenvalue equation are

m` () =

Chap. 3 = eigenfunctions of b
Lz and b
L 2 are

Y`,m` (, ) = `,m` ()m` ()

Chap. 4 = fully separated form of a SAME is

E,`, (r) = RE (r)`,m` ()m` ()

1
2

ei m` ,

for m` = an integer

Table 4.6. The development of the stationary angular momentum eigenstates for a rotationally
invariant 3D system.

The Laplacian in Cartesian and spherical coordinates. The 3D Laplacian in Cartesian coordinates
and spherical coordinates are
2
2
2
+
+
x2
y 2
z 2

1 2
2
2
b
sin
+
,
L = ~
sin

sin2 2

b2
1
L (, )
2
2
= = 2
r
+
.
r r
r
~2 r2
2 =

JQPMaster

Version: 8.35

(4.9.5a)
(4.9.5b)
(4.9.5c)

Printed: August 17, 2010

307
Bound states
boundary conditions

un,` (0) = 0
un,` (r) 0
r

limiting behavior

un,` (r) r`+1


r0

un,` (r) en,` r


r

physical admissibility

un,` (r) must be continuous for all r


un,` (r) must be smoothly varying for all r (continuous first
derivative dun,` / dr)
un,` (r) must be single-valued
un,` (r) must be normalizable:

b
Hermiticity of H

R
0

u2n,` (r) dr = 1

{ un,` (r) } must be orthogonal with respect to n:


R
un 0 ,` (r) un,` (r) dr = n 0 ,n
0
number of nodes in un,` (r) is nr = n ` 1

Table 4.7. Properties of the reduced radial function for any spherically symmetric potential
energy. Using these constraints and the classification of regions in Tbl. 4.5, you can draw accurate
rad
qualitative sketches of Pn,`
(r) for SAMEs (see footnote 39).
r`+1

Nn,`
normalization
constant

dominant factor as
r 0; ensures that
un,` (0) = 0
so n,`,m` (r) will be
physically admissible

fn,` (r)

en,` r

a finite polynomial of degree


nr = n ` 1 that controls
intermediate behavior
between the limits and the
number of nodes nr .

dominant factor as
r ; ensures bound
state behavior.

Table 4.8. The mathematical structure of the reduced radial functions for a particle in
a central potential energy.

Bound states

Continuum states

value at r = 0

un,` (0) = 0

uE,` (0) = 0

behavior in the limit r

un,` (r) 0

|uE,` (r)| <

energies

En,` < 0 quantized

E > 0 continuous

energy degeneracy

g(En,` ) 2` + 1
R 2
un,` (r) dr = 1
0

-fold degenerate

normalization

Table 4.9. Key properties of the reduced radial functions for bound and continuum stationary
angular momentum eigenstates. Bound state radial functions are labeled by the principal quantum number
n, continuum state functions by the (continuous) energy E > 0. (The normalization condition for continuum
states in the last row is included for completeness. These functions can be multiplied by any r-independent
factor and still satisfy the continuum radial equation; the choice /2 conforms to the conventions used in this
book.) We take bound-state wave functions to be real.

The radial and angular kinetic energy operators.

JQPMaster

Version: 8.35

The 3D kinetic-energy operator is

Printed: August 17, 2010

308

4.9. Users guide to Chap. 4

CF region including r = 0
CA region
CF region including r

L
0 r rn,`

un,` (r) rises from zero, varying in


the limit of small r like r`+1

R
L
r rn,`
rn,`

un,` (r) oscillates, though not with


a constant wavelength

R
rn,`
r<

un,` (r) decays to zero and for large


enough r assumes the form en,` r

Table 4.10. Qualitative behavior of the bound state radial function for a continuous spherically symmetric potential V (r) in classically allowed (CA) and classically
forbidden (CF) regions. The limiting behaviors for r 0 and r apply only in the
indicated limits, not throughout their respective regions.

in terms of Rn,` (r)


R

normalization integral

orthonormality integral

rad
(r)
radial probability density Pn,`

mean radius hrin,`

2
Rn,`
(r) r2 dr

Rn0 ,` (r)Rn,` (r) r2 dr

in terms of un,` (r)


R
0

R
0

u2n,` (r) dr
un0 ,` (r)un,` (r) dr

2
(r)
r2 Rn,`

u2n,` (r)

2
r Rn,`
(r) r2 dr

u2n,` (r) r dr

Table 4.11. Orthonormality and physical properties in terms of the radial and
reduced radial functions.

1 b2
r2
+
L (, )
r
2mr2

2
2
d
2 d
b rad (r) = ~ 1 d r2 d = ~
where T
+
2m r2 dr
dr
2m dr2
r dr

2
b =T
b rad (r) + T
b ang = ~ 1
T
2m r2 r

and

b ang =
T

1 b2
L
2mr2

(4.9.6a)
(4.9.6b)
(4.9.6c)

b rad + T
b ang + V (r). Only the potential
b=T
The Hamiltonian for a spherically symmetric potential V (r) is H
energy and the mass in the kinetic-energy operator are system-dependent.
Orthonormality relations for SAMEs. Much of the analysis of central-force problems follows from orthonormality conditions and their consequences for physically significant quantities like the radial probability
density (see Tbl. 4.11).

JQPMaster

Version: 8.35

Printed: August 17, 2010

309

4.10. Selected readings & references for Chap. 4


Employ your time in improving yourself
by other mens writing so that
you shall come easily by
what others have labored hard for.
Socrates

4.10.1 Numerical solution of the Schr


odinger equation
Libraries are full of useful books on numerical solution of differential equations.
The following books contain succinct discussions of relevant physics and guidance on how to work problems on
your PC, including how to solve the reduced radial equation.
(1) Theoretical Physics on the Personal Computer by E. W. Schmid, G. Spitz, and W. L
osch (Berlin: SpringerVerlag, 1997). Chap. 4 covers numerical integration and Chap. 13 the solution of the radial equation. See also
Chap. 17 on spherical harmonics.
(2) Quantum Mechanics on the Personal Computer, Third Edition, by S. Brandt and H. D. Dahmen (Berlin:
Springer-Verlag, 1994), especially Chap. 7 on the central force problem and Chap. 9 on special functions.
The following general texts on computational physics contain material relevant to this chapter.
(3) Computation in Modern Physics by William R. Gibbs (Singapore: World Scientific, 1994).
(4) Computational Physics by Nicholas J. Giordano (Upper Saddle River, NJ: Prentice-Hall, 1997).
(5) Computational Methods in Physics and Engineering by Samuel S. M. Wong (Englewood Cliffs, NJ: PrenticeHall, 1992). Chapter 8 covers numerical approaches to ordinary and partial differential equations.
If you want to learn how to solve the Schr
odinger equation (and other differential equations) using Mathematica,
herein order of increasing difficultyare some books that will help you:
(6) Elementary Numerical Computing with Mathematica , by Robert D. Skeel and Jerry B. Keiper (New York:
McGraw Hill, 1993).
(7) A First Course in Computational Physics by Paul L. DeVries (New York: Wiley, 1994). See especially Chaps. 5
and 7.
(8) Computational Physics: Problem Solving with Computers by Rubin H. Landau and Manuel J. P
aez (New
York: Wiley 1997).
(9) Chapter 4 of Mathematica in Theoretical Physics by Gerd Baumann (New-York: Springer-Verlag, 1996) is
devoted to quantum mechanical applications (including the Schr
odinger equation) as is Chap. 6 of Mathematica
for Physicists by Robert L. Zimmerman and Fredrick I. Olness (Reading, Mass.: Addison-Wesley, 1995). The
latter book includes a variety of commands you can use when solving the radial equation. An especially good
treatment of shooting methods appears in Chap. 10 of Mathematica for Scientists and Engineers by Thomas
B. Badher (Reading, Mass.: Addison-Wesley, 1995).
(10) Numerical Solutions for Partial Differential Equations: Problem Solving Using Mathematica by Victor
G. Ganzha and Evgenii V. Vorozhtsov (New York :CRC Press, 1996) contains a variety of sophisticated general
approaches to solving PDEs.
odinger equation is the Numerov al(11) One of the most powerful (and popular) methods for solving the Schr
gorithm. Youll find a good introduction to this approach in J. M. Blatt. Practical points concerning the
solution of the Schr
odinger equation. J. Comput. Phys., 1:382396, 1967.
(12) In one of their more colorful feats of nomenclature, numerical analysts have given the name shooting methods
to the class of techniques that search for eigenvalues by propagating the solution of a differential equation to
large values of the independent variable where it must satisfy prescribed boundary conditions. The literature
on these powerful methods is vast. A useful contextualization of the Numerov method in the class of shooting
methods appears in Shooting methods for the Schr
odinger equation by J. Killingbeck, Journal of Physics B,

JQPMaster

Version: 8.35

Printed: August 17, 2010

310

4.11. Exercises & problems for Chap. 4


20, 1411, (1987). Special methods that simplify such techniques when the potential being shot is symmetric are
discussed by H. Kobeissi, A. El-Hajj, and M. Kobeissi, Journal of Physics B, 23, 5725(1990). And youll find
a FORTRAN program you can download described in SSM: a set of subprograms for calculating eigenvalues
for a diatomic molecule using a simplified shooting method. by H. Kobeissi, A. El-Hajj, and M. Kobeissi,
Computer Physics Communications, 74, 297 (1993).

(13) To learn about how to improve your numerical answers with very little work, see Eigenvalue Correction
Estimates for the One-dimensional Schr
odinger Equation by James D. Talman, Journal of Computational
Physics 37, 19, (1980). In particular, a dramatic improvement on the Numerov algorithm is the renormalized
Numerov method developed by B. R. Johnson and described in his article in Journal of Chemical Physics
67, 4086, (1977). See also the articles on estimating errors by A. D. Paptis and J. R. Cash, Computer Physics
Communications 36, 113, (1985), on how to improve the shooting method by F. Y. Hajj, Journal of Physics
B 13, 4521, (1980) and, finally, on a very efficient matrix formulation of this method by B. Lindberg, American
Journal of Physics 88, 3805, (1988).

4.10.2 Programs for solving the radial equation


Many physicists and chemists have generously made available their programs to solve various formidable quantum
problems. A host of these tackle the radial equation (which many authors call the one-dimensional Schr
odinger
equation). Here are a couple I particularly recommend:
odinger Equation Using
(1) A FORTRAN Program for the Numerical Integration of the One-Dimensional Schr
Exponential and Bessel Fitting Methods, by J. R. Cash, A. D. Raptis, and T. E. Simos, Computer Physics
Communications 56, 391, (1990).
(2) An Integral Equation Program to Calculate Radial Wave Functions and Scattering Phase Shifts of ShortRange Local Interactions, by M. S. Stern, Computer Physics Communications 17, 365, (1979). This
program concerns continuum (scattering) states and transforms the TISE and its boundary conditions into a
single integral equation.

4.11. Exercises & problems for Chap. 4


No matter how big and tough a problem may be,
get rid of confusion by taking one little step towards solution.
Do something.
George F. Nordenholt

I Thought and Review Questions


Q 4.1.

The deuteron: qualitative perspectives I


In the historical development of nuclear physics, the deuteron, which consists of one proton and one neutron, was
instrumental in understanding nuclear structure and the properties of nuclei. To an excellent approximation, this twonucleon system can be modeled as a single pseudo-particle whose mass is the reduced mass of the proton and
neutron. In this model, the particle experiences a central, attractive potential energy V (r). Experimental measurements have shown that the deuteron has one and only one bound state of total energy En,` = 2.226 0.003 MeV,
where 1 MeV = 106 eV (Appendix L).
Since no one knows an explicit closed form for the potential energy for the deuteron, it must be modeled (see
The Principle of Modeling in Appendix Q). One of the most accurate such model potential energies is the Yukawa
function (Complement 11.5), which in this problem Ill write as
V (Y) (r) V0

er/r0
,
r/r0

Yukawa potential energy.

(4.1.1)

The Yukawa potential energy depends on two parameters which control its range and strength: r0 and V0 . (The
Yukawa potential energy is nearly zero for r > 2r0 . So r0 is a sensible estimate of the range of the deuteron potential
energy.) For the deuteron these parameters are determined from experimental data to be

JQPMaster

Version: 8.35

Printed: August 17, 2010

311
r0 = 2.0 fm = 2.0 1015 m

range parameter

(4.1.2a)

V0 = 36.0 MeV = 36.0 106 eV

well depth,

(4.1.2b)

where 1 fm = 1 fermi = 1 1015 m is the characteristic length of a nucleus. Fig. 4.1.1 shows the Yukawa function
and the bound-state energy of the deuteron.49

Figure 4.1.1. The Yukawa model


of the potential energy of the
deuteron. The horizontal dashed line
shows the experimentally determined
energy of the single bound state supported by this potential energy. The
units of distance are fermi; the units
of potential energy are MeV.

potential energy HMeVL

2
0
-2
-4
-6
-8
0

r HfmL

(a) What is the complete set of commuting operators (CSCO) for the deuteron in this model? List and identify
the quantum numbers that label the bound stationary angular momentum eigenstates (SAMEs) of the system.
(b) Is the z projection of the orbital angular momentum in this bound state sharp? If so, explain why and give its
value. If not, explain why not.
(c) Does this state have definite parity? If so, what is its parity? If not, why not?
(d) What are the boundary conditions on the reduced radial function for this state?
(e) How many nodes does the reduced radial function for this state have? Why?
(f) How many classical turning points does this reduced radial function have? Estimate the value(s) of the
classical turning point(s) in units of fermi.
(g) On a sketch of the deuteron potential energy, show and label the classical turning points and the classically
allowed and classically forbidden regions for this bound state. Also draw a careful qualitative sketch of un,` (r).
(h) The Yukawa potential energy supports continuum states as well as bound states. In what way(s) would the
reduced radial function for a continuum state of the same orbital angular momentum differ from the bound
state you drew in part (g)?
(i) Experimental evidence indicates that the deuteron has a nonzero electric quadrupole moment. Discuss the
implications of this finding for the Yukawa model. Hint: Think about the orbital angular momentum properties
of the state.
Q 4.2.

Screening via the Yukawa potential energy


In atomic physics the Yukawa potential energy is often used to model quite complicated phenomena. In a manyelectron atom, the valence electron does not experience the full attractive Coulomb force exerted by the nucleus. This
attractive force is screened by the other electrons (see Complement 4.D and Chap. 11). The combined effect of
the electron-nuclear attraction and screening is that the potential energy of the valence electron is weaker than the
Coulomb potential energy due to the nucleus alone. This phenomenon can be modeled by a Yukawa function that
depends on a screening parameter ; in atomic units (Appendix F), this function is
V (Y) (r) =

Z r/
e
,
r

(4.2.1)

where Z is the nuclear charge, and e02 1 in atomic units.


49 Read on: For more work with this model, see Question 4.2 and Exercise 4.8. For alternative models of the nuclear potential
energy, see Question 4.3, Exercise 4.9 and Problems 4.14.

JQPMaster

Version: 8.35

Printed: August 17, 2010

312

4.11. Exercises & problems for Chap. 4

Suppose the screening parameter is large enough that Eq. (4.2.1) supports several bound states. Specifically,
consider = 1500 a0 . Figure 4.2.1 shows reduced radial functions un,` (r) for two bound d states of this potential
energy. One of these states has energy En,` = 2.99 eV; the other has energy 5.17 eV.
u n d HrL

state a

Figure 4.2.1.
Reduced
radial
functions
for
two bound states of the
Yukawa potential energy.
In Eq. (4.2.1), the screening
parameter is = 1500 a0 .
Both are d states.

state b
r

(a) What are the values of the quantum numbers n and ` for these radial functions?
(b) Which of these states has the largest binding energy? Why did you pick this state?
(c) For each state, write down the analytic dependence of the radial probability density in the limit r 0.
(d) For each state, write down the analytic dependence of the radial probability density in the limit r .
max
Table 4.2.1 contains the mean values hrin,` of r and the values rn,`
at which the radial probability densities for these
states attain their maximum values. Lets look first at state a, whose reduced radial function appears in Fig. 4.2.2
along with the mean radius (dashed vertical line) and the radius of maximum probability density (solid vertical line).
max
(e) According to the table, for state a, rn,`
< hrin,` . Why? Hint: Consider the explicit definitions of the mean
radius and the radial probability density.

(f) The numbers in Tbl. 4.2.1 reveal an interesting difference between the two states:
max
rn,`
< hrin,`

for state a,

(4.2.2a)

max
rn,`

for state b.

(4.2.2b)

> hrin,`

Explain this difference. Hint: Consider the effective potential energy and the radial probability densities for
these states.
Table 4.2.1. The mean radius
hrin,` and the radius of maximum radial probability denmax
sity rn,`
for the states in
Fig. 4.2.1. All values are in
atomic units (Appendix F).

Q 4.3.

state
state a
state b

max
rn,`
(a0 )

9.0
78.8

hrin,` (a0 )
10.75
70.75

The shell model of the nucleus


Nuclei typically contain several nucleons (protons and neutrons). Early studies of the physics of nuclei made
extensive use of the shell model (Chap. 12), in which each nucleon experiences the same single-particle spherically
symmetric potential energy V (r). This model enables the study of nuclei in terms of solutions of the single-particle
Schr
odinger equation for a nucleon of mass m,

~2 2

+ V (r) E E (r) = 0.
(4.3.1)
2m
Although the Yukawa potential energy was most often used for the nuclear potential energy in such studies (see Question 4.1, another highly successful model is the potential energy of a three-dimensional isotropic simple harJQPMaster

Version: 8.35

Printed: August 17, 2010

313
u n d HrL
0.4

Figure 4.2.2. The reduced radial functions for state a of


Fig. 4.2.1. This is one of the
bound d states of the Yukawa function Eq. (4.2.1) with screening parameter = 1500 a0 . Also shown
are the mean radius (dashed vertical line) and the radius of maximum probability density (solid
vertical line) for this state (see
Tbl. 4.2.1).

state a

0.3

0.2

0.1

10

20

30

40

monic oscillator (3D SHO):

1
m 2 r2 .
2
For nuclei, the frequency is related to the nuclear mass number A by ~ 41 A1/3 MeV.
V (r) =

(4.3.2)

(a) According to this model, what observables are constants of the motion of the nucleus?
(b) Do there exist separable bound-state Hamiltonian eigenfunctions for this potential energy? Why or why not?
(c) Are all Hamiltonian eigenfunctions for this potential energy of separable form? Why or why not?
(d) Do the stationary state energies depend on m` ? Why or why not?
(e) Write down the analytic dependence of the bound-state reduced radial functions un,` (r) in the limit r 0.
(f) Write down the analytic dependence of these functions in the limit r .
(g) Determine the energy eigenvalues of the first three s states of the 3D SHO, writing your answer in terms of
~ and . Hint: Think; dont calculate. You need not perform any analysis or algebra to solve this problem.
(h) Evaluate the ground-state energy (in MeV) for the 3 He and 4 He isotopes of helium. (The pre-superscript is
the mass number A.)
(i) Solution of the radial equation yields not just s-state energies but also En,` for ` 0. Some of these energies
turn out to be degenerate: for example, E2,0 = E1,2 . In fact, the energies of the first four energy levels are
related by
ground state
first excited state
second excited state
third excited state

E1,0
E1,1
E2,0 = E1,2
E2,1 = E1,3

According to the shell model, nuclei are constructed by filling these single-particle energy levels a nucleon at a
time. As well discover in Chap. 9, we can assign two (and only two) nucleons of a given type to each stationary
angular momentum eigenstate (SAME). Once we have allocated nuclei to the SAMEs that correspond to a
given energy level, we say that the shell defined by that energy is filled or closed.50 Based on the 3D SHO
model, how many nucleons must a nucleus contain to fill the (1, 0) shell? Also answer this question for shells
with n = 2, 3, and 4.
Q 4.4.

The structure of Hamiltonian eigenfunctions.


Beginning with the time-independent Schr
odinger equation for a spherically symmetric potential energy V (r), isolate
L 2 n,`,m` (r); that is, write this equation in the form
the term b
b
L 2 n,`,m` (r) = stuff.

(4.4.1)

50 Commentary: The reason for this restriction is that nucleons have spin 1/ , with projection on the axis of quantization
2
of 1/2 (Chap. 6). Two nucleons can have the same quantum numbers (n, `, m` ) only if they differ in their spin projection
quantum number.

JQPMaster

Version: 8.35

Printed: August 17, 2010

314

4.11. Exercises & problems for Chap. 4

Using only your equation, explain why n,`,m` (r) must be the product of a radial function times a spherical harmonic.
Q 4.5.

The interplay of math and physics in the radial function


The effective potential energy V`eff (r) imposes a lower bound on the bound-state energies for any state with ` > 0.
That is, there are no physically admissible SAMEs for energies E < V`eff .
(a) Consider a fixed value of ` > 0. Using only the radial equation for a potential energy V (r), show that if
E < V`eff , then the solution uE,` (r) cannot satisfy the boundary conditions required for a bound state.
(b) Using only your answer to part (a), explain why all bound-state radial functions un,` (r) must be oscillatory
over some finite domain of r.
(c) At what value of r does V`eff (r) for a d state (` = 2) attain its minimum? What is the value of V`eff (r) at this
value of r? (Give your answers in atomic units.) Using your answers and the expression for the stationary-state
energies of atomic hydrogen in atomic units, En = (1/2n2 ) E h (Chap. 5), explain why there do not exist
bound one-electron d states with principal quantum numbers n = 1 or n = 2.

I Exercises
E 4.1.

The Three-Dimensional Oscillator


(a) What are the allowed values of the principal quantum number n for a one-dimensional harmonic oscillator?
What are the energy and parity of the nth stationary state?
(b) The Hamiltonian eigenfunctions for a particle of mass m in a three-dimensional potential energy well,
V (r) = m 2 r2 /2, can be written as the separable product of functions in x, y, and z with respective quantum
numbers nx , ny , and nz . Determine the energy, parity, and degree of degeneracy of the lowest four energy
levels. For each level, list all possible sets of quantum numbers (nx , ny , nz ).
(c) Were you to solve the time-independent Schr
odinger equation for the three-dimensional oscillator in spherical
coordinates, you would find the same energy eigenvalues you found part (b) but different eigenfunctions. Using
your knowledge of the parity and degeneracy of various states in any central potential energy, determine
which values of the orbital quantum number ` are associated with each of the four energy levels of part (b).
(More than one value of ` may be associated with a given energy.)

E 4.2.

Nodes and antinodes


The nodes of a function f (r) are values of r in the open interval 0 < r < where the function is zero. The antinodes
of f (r) are the values of r where the function attains a local maximum. How many antinodes does un,` (r) for a bound
state in a spherically symmetric potential energy have? Justify your answer with a argument and/or sketches.51

E 4.3.

The radial probability density for a central potential energy


If a potential energy is highly symmetric, we can often gain insight into the bound states it supports even if we know
very little about the states themselves. For example, consider a single particle in a spherically symmetric potential
energy whose sole quantum state is some eigenstate of b
L 2 and b
Lz . What can you conclude about the dependence
of the position probability density for such a state? What does your conclusion imply about the symmetry of this
function about the z axis? Prove or otherwise justify your answers.

E 4.4.

Central versus non-central forces


Although spherically symmetric potential energies are extremely important, most potential energies in nature are not
spherically symmetric. So its important to understand whats special about central-force systems and to think a
little about how to handle systems with lower symmetry.
b of this system
(a) For a system with a spherically-symmetric potential energy V (r), prove that the Hamiltonian H
2
b
b
commutes with L and Lz .
51 Commentary: This information is useful in solutions of the reduced radial equation by numerical integration. In such
approaches, one often counts nodes or antinodes to determine the principal quantum number n for a numerically generated
solution un,` (r).

JQPMaster

Version: 8.35

Printed: August 17, 2010

315
(b) Now suppose the potential energy depends on but not on , as V (r) = V (r, ). For such a potential energy,
answer the following questions, justifying each answer:52 (i) Does the Hamiltonian of this system commute
with b
L 2 ? (ii) Does the Hamiltonian commute with b
Lz ? (iii) Does there exist a complete set of simultaneous
b b
eigenfunctions of {H,
L 2, b
Lz }?
E 4.5.

Separability in special cases.


There is a special class of potential energies V (r) for which the TISE is separable even though this function depends
on and . Suppose V (r) has the form
V (r, , ) = Vr (r) +

1
1
V () + 2 2 V (),
r2
r sin

(4.5.1)

where each factor Vr (r), V (), and V () depend only on the variable indicated in its argument.
(a) Do solutions (r) = f (r) g() h() to the TISE exist for such a potential energy of the form (4.5.1)? Prove
your answer.
(b) If your answer to part (a) was yes, derive equations for the factors f (r), g(), and h().
(c) Under what conditions on V () is h() equal to an eigenfunction m` () of b
Lz ?
(d) Under what conditions on V () is the product g()h() equal to a spherical harmonic Y`,m` (, )? Must V ()
satisfy the condition in part (c) for this separability to hold? Why or why not?
E 4.6.

The indicial equation.


(a) By examining what happens when you insert the expansion Eq. (4.6.18), p. 295 into the radial equation and
take the limit r 0, explain why only the leading term rs is relevant in this limit.
(b) Use the expansion Eq. (4.6.18), p. 295 to derive an equation (called the indicial equation) for s which
prescribes that either s = ` and s = ` + 1.

E 4.7.

Mysteries near zero.


As noted in 4.6, mathematics provides two solutions to the near-zero radial equation (4.6.14): one proportional to
r`+1 , the other to r` .
(a) By examining the radial probability density in the near-zero limit, show mathematically why we must reject
one of these solutions.
(b) Argue from the Born interpretation of the radial probability density that a wave function whose radial factor un,` (r) behaves as r` in this limit cannot represent a bound state.

E 4.8.

Yukawa potential energy sketches.


Figure 4.1 compares the Yukawa potential energy Eq. (4.1.1), p. 310 for A = 10.82 and = 1 to the pure-Coulomb
potential energy [Eq. (4.2.5), p. 269] for Z = 1. Sketch the ground and first excited s states of this Yukawa potential
energy. Also sketch the lowest-lying p and d states.

E 4.9.

Woods-Saxon potential energy sketches.


The Yukawa potential energy of Question 4.1 is by no means the only useful model of the nuclear potential energy.
An alternative is the Woods-Saxon function,
V (r) =

V0
,
1 + e(rR)/

Woods-Saxon potential energy.

(4.9.1)

52 Commentary: The non-central potential energy considered in this problem occurs in many real systems. For example, the
potential energy that describes the interaction of an electron with a diatomic molecule depends on but not on because a
diatomic molecule has axial symmetry . Such potential energies also arise in the study of magnetic properties of materials.

JQPMaster

Version: 8.35

Printed: August 17, 2010

316

4.11. Exercises & problems for Chap. 4


0.0

Yukawa

-0.2

pure Coulomb HZ=1L


-0.4
VHrL
-0.6

-0.8

Figure 4.1.
The pure-Coulomb and
Yukawa Potentials. Both potential energies are in atomic units (Appendix F).

-1.0

10

This function has three parameters: a well depth V0 ; a measure R of the range of V (r); and , which controls how
rapidly the function goes to zero with increasing r.
(a) Graph the Woods-Saxon function for several choices of the parameters. Try enough choices to get a feeling
for the effect of varying these parameters on the shape of V (r).
(b) Graph the Woods-Saxon function for V0 = 8 MeV, R = 5.17 fermi, and = 0.75. On the same graph, show
a pure-Coulomb potential energy [Eq. (4.2.5), p. 269] for Z = 1. On this graph, sketch the ground and first
excited s states of both potential energies.
E 4.10.

An equation for the auxiliary function fn,` (r).


Derive the following differential equation for the intermediate-r function fn,` (r) in Eq. (11.3.5a), p. 685:

d2 f
`+1
df
`+1
+
2

U
(r)
+
2

f (r) = 0.
dr2
r
dr
r

(4.10.1)

This equation is useful because it can be solved via the method of power series.

E 4.11.

Thinking about bound-state energies.


(a) Write down the radial equations for ` = 0 and for ` = 1 for a potential energy V (r). Explain why, for an
arbitrary attractive V (r), the energies En,` and the minimum-energy reduced radial functions for ` = 0 and
` = 1 must be different. Support your argument with sketches of V`eff (r) and un,` (r).
(b) Suppose V (r) supports one ` = 0 bound statethat is, there exists a normalizable radial function R1,0 (r) for
energy E1,0 < 0. Does it necessarily follow that V (r) will support at least one bound state for all ` > 0? Does
it follow that V (r) will support any bound states for ` > 0? Explain your answers.

E 4.12.

A gaggle of mystery states.


For the states represented (at time t = 0) by each of the following initial wave functions, identify the constants of
motion, state whether or not the states are stationary, and state whether or not they are SAMEs. In each case,
justify your answers.
1 (r, 0) = n,2,0 (r, , ) + 2n,2,1 (r, , ),
2 (r, 0) = 3,1,0 (r, , ) + 52,1,1 (r, , ).

JQPMaster

Version: 8.35

Printed: August 17, 2010

317

I Problems
P 4.1.

** Spherically Symmetric Quantum States


Although in general, spherical symmetry of a potential energy does not imply spherical symmetry of its Hamiltonian
eigenfunctions, there are special states whose Hamiltonian eigenfunctions depend only on r: for these spherically
symmetric states the eigenfunction is E (r) = E (r), independent of and .
(a) Show that for spherically symmetric states the magnitude and z projection of the orbital angular momentum
are zero.
(b) Your work on part (a) implies that the effect of the Laplacian on such states is especially simple:
"
#

1 d
1 d2
2
2 d
E (r) = 2
r
E (r) =
rE (r) .
(for spherical states only)
r dr
dr
r dr2

(4.1.1)

The second equality suggests that we might simplify the TISE for E (r) by getting rid of a factor of r. Derive
the equation for reduced radial function uE (r) in E (r) uE (r)/r.
To illustrate the properties of spherical states, lets turn to the hydrogen atoman electron and a proton interacting
via Coulomb forces. As well see in Chap. 5, we can reduce this system to the relative motion of a particle of mass
(the reduced mass of the atom) moving in the pure-Coulomb potential energy [Eq. (4.2.5), p. 269].
(c) Sketch un,0 (r) for the lowest three spherical bound states of the pure-Coulomb potential energy (n = 1, 2,
and 3). Explain how you deduced the form of these functions.
(d) Now focus on the ground state. First, show that the ground-state reduced radial function must decay exponentially as r becomes infinite:
u1,0 (r) er ,
(4.1.2a)
r

where 2 2E1,0 /~2 . Explain why this implies that u1,0 (r) must have the form
u1,0 (r) = f (r) er ,

(4.1.2b)

where f (r) in Eq. (4.1.2b) must be a finite polynomial that approaches 0 as r 0.


(e) The simplest such polynomial is f (r) = r. Show that plugging f (r) = r into Eq. (4.1.2b) solves the equation
for u1,0 (r) you derived in part (b). Determine the corresponding eigenvalue E1,0 . Show that the constant
is the inverse of the first Bohr radius a0 ~2 /me e02 .
(f) Normalize u1,0 (r). Now graph u1,0 (r) and the corresponding ground-state eigenfunction 1,0,0 (r). Hint:
For technical reasons, youll find it easier to plot 1,0,0 (r)/1,0,0 (0) and u1,0 (r)/[a0 u1,0 (0)].
(g) In a spherically symmetric state, the probability of detecting the particle in a spherical shell of thickness dr
at r is the radial probability density times the appropriate volume element, 4|1,0,0 (r)|2 r2 dr. Graph the
(normalized) probability density for the ground state. What is the expectation value of r for this state? What
is the most probable value of r for detecting the particle?

P 4.2.

* Exploring a reduced radial function and its radial probability density.


Figure 4.2.1 shows the effective potential energy and the reduced radial function for a SAME of a spherically symmetric
potential energy V (r).
(a) What can you conclude about the relationship between n and ` for this state?
(b) On the right panel in Fig. 4.2.1, sketch the radial probability density for this state as accurately as you can.
(c) Consider a spherical shell of thickness r = 0.1 a0 and radius r0 . For what value of r0 is the particle most
likely to be found in such a shell?
(d) The mean radius in this state is 10.75 a0 . Is this value larger, smaller, or equal to the shell radius r0 of part (c)?
Explain your answer, using your sketch of the radial probability density in part (b).

JQPMaster

Version: 8.35

Printed: August 17, 2010

318

4.11. Exercises & problems for Chap. 4


0.4

0.15

0.3

0.10
0.05

0.2

0.00

u n,{

0.1

P n,{
-0.05

0.0
-0.10
-0.1
-0.2

-0.15
0

10

15

20

25

30

-0.20

10

r Ha0 L

15

20

25

30

r Ha0 L

Figure 4.2.1. Left panel: an effective potential energy (solid line) and a reduced radial function (dotted line). The
dashed horizontal line shows the bound-state energy for this state. Right panel: Your radial probability density for
Problem 4.2.

P 4.3.

* Comparing radial functions.


Figure 4.3.1 shows a second radial function (the solid curve) for the same quantum number n as the state in Problem 4.2 (the dashed curve in Fig. 4.3.1).
(a) Is ` for the state in Fig. 4.3.1 larger, smaller, or equal to ` for the state in Fig. 4.2.1? Why?
(b) Consider a sphere of radius 5 a0 . For which of the states in Fig. 4.3.1 is the probability of finding the electron
anywhere inside this sphere the largest? What feature of the reduced radial function of this state justifies your
answer?
(c) Suppose instead that we consider a sphere of radius 1 a0 . Is your answer to part (b) the same or different? Why?
(d) Is the average value of the radial position of the electron larger or smaller for the state in Fig. 4.3.1 than for
the state in Fig. 4.2.1? Why?
0.4

State 1
0.2

u n,{

Figure 4.3.1. The reduced radial


functions for Problem 4.3. The
dotted curve labeled State 1 is the
state discussed in Problem 4.2 and
shown in the left panel of Fig. 4.2.1.

P 4.4.

0.0

State 2

-0.2

-0.4

10

15

20

25

30

r Ha0 L

* A tale of three radial functions


Figure 4.4.1 shows the states of Problems 4.2 and 4.3 along with yet a third state for the same n. For this potential
energy, these are the only bound radial functions for this n.
(a) What is the value of ` for each of the three states?
(b) What is the value of n for these states?
(c) For which state is the electron most likely to be found inside a sphere of radius r0 = 1 a0 ? For which state is
it most likely to be found inside a sphere of radius r0 = 5 a0 ? If your answers are the same, explain why; if
theyre different, explain which features of the radial function are responsible for the difference.
(d) Write down the explicit functional dependence of the limiting forms of these three radial functions as r 0.
JQPMaster

Version: 8.35

Printed: August 17, 2010

319
(e) For which state is the mean radius the largest? Why?
(f) For which state is the uncertainty in r the largest? Why?
0.4

State 1

State 3

0.2

u n,{

0.0

State 2
-0.2

Figure 4.4.1. The reduced radial functions for


Problem 4.4.
The states labeled state 1 and
state 2 are those of Problem 4.2 and Problem 4.3,
respectively. See Figs. 4.2.1 and 4.3.1.

P 4.5.

-0.4

10

15

20

25

30

r Ha0 L

* Variation of the radial function with principal quantum number.


Figure 4.5.1 shows radial probability densities for three stationary angular momentum eigenstates with the same `.
(a) For which of these statesa, b, or cis the principal quantum number n the largest? Why?
(b) For which state is the binding energy n,` the largest? Why?
(c) For which state is the uncertainty in the radial position the largest? Why?
(d) For which state is the mean radial position the smallest? Why?
(e) For each state, what is the functional dependence of Pn,` (r) as r 0?
(f) Is ` = 0 for these states? Why or why not?
0.12

state a

0.10
0.08
P n,{ 0.06

state b

0.04

state c

0.02

Figure 4.5.1. Three radial probability densities for Problem 4.5.


Each of these three states has the
same orbital quantum number `.

0.00
0

10

20

30

40

50

60

r Ha0 L

P 4.6.

* Variation of the radial function with orbital angular momentum quantum number.
The solid curve in the Fig. 4.6.1 shows the effective potential energy for a p state. The dotted curve shows un,` (r) for
an s state with principal quantum number n. On this figure, sketch un,` (r) for the p state with this n. Try to make
your sketch as accurate as possible, taking account of the variation with r of the local wavelength and amplitude
(4.7). Assume thatthe binding energies do not depend significantly on `.

P 4.7.

* Further investigation of the effect of the principal quantum number.


The solid curve in Fig. 4.7.1 shows un,` (r) for a state with ` = 3 and n = 4. On this figure, sketch un,` (r) for ` = 3
and n = 5. Assume that the binding energies for this system increase with increasing n. Hint: Remember that both
reduced radial functions must satisfy by physically admissible.

JQPMaster

Version: 8.35

Printed: August 17, 2010

320

4.11. Exercises & problems for Chap. 4

0.4

0.2

u n,{

Figure 4.6.1. The effective potential


energy for a p state (thick solid curve)
and the reduced radial function for an
s state (dotted curve). The dashed horizontal line shows the bound-state energy for
the p state.

0.0

-0.2

-0.4
0

10

12

14

r Ha0 L
0.4
0.3
0.2
0.1
u n,{
0.0
-0.1

Figure 4.7.1. The effective potential


energy (thick curve) and the reduced
radial function (dotted curve) for a
state with ` = 3 and n = 4.

P 4.8.

-0.2
-0.3

10

20

30

40

50

60

r Ha0 L

** Asymptotic behavior of the radial probability density


From the generic behavior of n,`,m` (r) for an atom in any model whose potential energy is spherically symmetric,
rad
show that Pn,`
(r) for large r is related to the first ionization energy I by53
rad
Pn,`
(r) ' A rn e2

2I r

(4.8.1)

where A is a constant and n is the principal quantum number.54


P 4.9.

* Guidance from the effective potential energy.


Figure 4.9.1 shows V`eff (r), the bound-state energy En,` , and un,` (r) for a particular state of a purely attractive
potential energy (the solid curve). Also shown are V`eff (r) and En,` for another state with the same n and the same `
for a second, stronger attractive potential energy (the dashed curves). On this figure, sketch un,` (r) for the stronger
potential energy.

P 4.10.

**** The radial kinetic-energy operator


A rather subtle aspect of 3D quantum mechanics is related to the kinetic-energy operator in spherical coordinates.
One of the first tactics you learned in classical physics is to separate the kinetic energy into a radial kinetic energy
and an angular kinetic energy . The radial term involves the component of the linear momentum p along the radial
unit vector er ; the second term involves the orbital angular momentum L = r p:
T =

p2r
L2
+
,
2m
2mr2

where

p r = er p =

r
p,
r

(classical kinetic energy operator).

(4.10.1)

53 Jargon: The first ionization energy is the amount of energy you must supply to an atom to eject the least-tightly-bound
electron so that it winds up at r with zero kinetic energy. See Chap. 10.
54 Commentary: This simple equation is the basis for a powerful approach to describing the asymptotic properties of wave
rad (r) in powers of 1/Z, where Z is the
functions for many-electron atoms. This approach is based on the expansion of ln Pn,`
atomic number.

JQPMaster

Version: 8.35

Printed: August 17, 2010

321
0.5

un { HrL

Potential 1

Figure 4.9.1. Two effective potential energies for Problem 4.9. The
solid curves show V`eff (r) and the radial function for the weaker of the
two physical potential energies V (r).
The dashed curves show V`eff (r) for
the stronger V (r). Both effective potential energies have the same `.

En {
Potential 2

-0.5

-1
0

10

15

20

r Ha0 L

In this chapter and in Chap. 3 we implemented a similar partitioning of T without mentioning the radial momentum:
using b
L 2 (, ) in spherical coordinates, we got

1 b2
~2 1
2
b
(r)
+
(4.10.2)
T(r)
=
r
L (r, , ) (r),
2m r2 r
r
2mr2
This approach, however, bypasses an interesting question: how can we generate components of the linear momentum
operator in coordinate systems other than Cartesian coordinates?
A first guess. From our familiarity with expressions for the Cartesian coordinates of b
p such as b
px = i ~ /x,
we might write down analogous expressions for the spherical components:
b
pr =

br

b
p = i ~
,
r
r

b
p = i ~

and

b
p = i ~

(4.10.3)

Although sensible, this approach leads to serious problems.


(a) Show that b
pr as defined in Eqs. (4.10.3) is not Hermitian.
This lack of Hermiticity arises from the volume element in spherical coordinates. Recall that in Cartesian coordinates,
d3 v = dx dy dz. In spherical coordinates, however, the volume element d3 v = r2 sin dr d d is the product of the
three differentials dr d d and a function f (r, ) = r2 sin . That function is the problem.
(b) We can eliminate the problem by getting rid of a certain asymmetry in the structure of the operator (4.10.3).
Show that the following symmetrized radial momentum operator is Hermitian:
b
pr =

1
2

br
br
b
p +b
p
r
r

symmetrized radial
momentum operator.

(4.10.4a)

(c) From
pr in spherical coordinates. First show that

this Hermitian operator, we can derive expressions for b


r, b
pr = i ~ . Then show how to write the operator (4.10.4a) as
b
pr =

br
1
b
p i ~ = i ~
r
r

1
+
r
r

(4.10.4b)

(d) Show that Eq. (5.C.4a) implies that


i
b
~2 1 2 h
p2r
rh(r)
.
h(r) =
2m
2m r r 2

(4.10.5)

(e) Finally, derive from Eq. (4.10.5) the quantum counterpart of the classical partitioning of energy Eq. (4.10.1)m
and identify the radial kinetic-energy operator .
p does not suffer from lack of Hermiticity.
(f) Show that the expression in Eq. (4.10.4a) for b

JQPMaster

Version: 8.35

Printed: August 17, 2010

322

4.11. Exercises & problems for Chap. 4

From the presence of along with r in the function f (r, ) that appears in the volume element, you might guess
that b
p = i ~ / is not Hermitian. You would be right. So we must construct a properly Hermitian operator
for b
p as we did for b
pr . Here is a general strategy you can apply to any component in any curvilinear coordinates
in 3D. You need only know the function that multiples the product of differentials in the volume element. For the
sake of generality, Ill denote the coordinates by q1 , q2 , and q3 and write f = f (q1 , q2 , q3 ). Then the corresponding
momentum is

[f (q1 , q2 , q3 )]1/2
qi

= i ~
+
f (q1 , q2 , q3 ) .
qi
2f (q1 , q2 , q3 ) qi

b
pi = i ~ [f (q1 , q2 , q3 )]1/2

(4.10.6a)
(4.10.6b)

(g) Show that the appropriately symmetrized (Hermitian) operator for b


p acts on a generic function of as55

i
1

1
h
b
p g() = i ~
sin g() = i ~
+ cot g().
(4.10.7)

2
sin
Remark: Underlying the trouble we ran into when converting to spherical coordinates is the lack of rotational
invariance of unit vectors in spherical coordinates. This makes even as writing down the gradient operator (let
alone the Laplacian) fraught with potential difficulties. The trick is knowing how to evaluate the partial derivatives
of the unit vectors that appear in the definition of the gradient in spherical coordinates,
= er

They are

1
1

+ e
+ e
,
r
r
r sin

er =

e ,

e = er ,

er =

e sin ,

e = e cos ,

e = (er sin + e cos ).

P 4.11.

e = 0,

(4.10.8a)
(4.10.8b)
(4.10.8c)

*** An electron interacting with an electric dipole .


The potential energy of an electron in the field of a fixed point dipole depends on the permanent dipole moment d
and the electron charge e > 0:
ed
V = 2 cos .
(4.11.1)
r
Hint: Use The Principle of Least Algebra throughout this problem! You may find it convenient to work in atomic
units, setting e 1 (see Appendix F).
Brainstorming questions.
(a) Does b
Lz belong to the CSCO for this system? Does b
L 2 belong to the CSCO? What angular momentum
observables of this system are conserved?
(b) Do you expect to be able to separate the dependence of the spatial functions on all three spherical coordinate
variables r, , ? Do you expect to be able to separate the dependence of the spatial functions on any of these
variables? Why or why not?
(c) Do there exist stationary states of definite parity? Are the Hamiltonian eigenfunctions proportional to spherical
harmonics? Are the Hamiltonian eigenfunctions proportional to m` ()? Why or why not?
(d) Are the Hamiltonian eigenfunctions proportional to `,m` ()? Why or why not?
55 Read on: See 5.6 of Introduction to the Quantum Theory (Third Edition) by David Park (McGraw-Hill, 1992)
for details concerning the derivation and application of this general expression. See also the article by J. D Hey in American
Journal of Physics, 61, 741 (1993).

JQPMaster

Version: 8.35

Printed: August 17, 2010

323
The problem.
(a) Graph the function (13.5.4a) and, on a separate figure, the Coulomb potential energy Eq. (4.2.5), p. 269.
Design your graphs to illustrate as clearly as possible the symmetry properties that distinguish these two potential
energies. These properties include (1) symmetry under inversion; and (2) rotational invariance.56
(b) Discuss each of your graphs briefly, pointing out the symmetry properties they reveal. (You may use different
graphs to illustrate different properties.)
Use your insight as guidance in considering the following, more analytical questions.
(c) Does b
L 2 commute with the dipole potential energy (13.5.4a)? Prove your answer by evaluating the commutator.
(d) Can you separate the dependence of (r) from this functions dependence on r and ? If not, explain, via
a mathematical proof or demonstration, why not. If so, do so: first write down the factor in (r), then
derive an equation for an auxiliary function F (r, ) that contains the r and dependence of (r).
(e) Try to separate the r and dependencies in the wave function. You may be surprised at what happens!
(f) Explicitly compare your equation for the dipole potential energy to the corresponding equation for a central
potential energy V (r); in particular, identify which terms are the same and which are different. (You will
need to do some minor algebraic manipulation to make your dipole equation look as much as possible like
the equation for a central potential energy.)
(g) Explicitly compare the radial equation for the dipole potential energy to the radial equation for a central
potential energy. Which terms are the same, and which are different?
(h) Does introduction of a reduced radial function un,` (r) into the radial equation for the electron-dipole system
result in the same simplifications that it does for a rotationally invariant system? Why or why not?
(i) Is the equation for the dipole potential energy satisfied by the function `,m` () we derived in Chap. 3? If
so, demonstrate that it works; if not, explain why not.
P 4.12.

** 1/2 Rotations of a diatomic molecule


In molecular physics, we use the quantum mechanics of angular momentum to understand rotations of diatomic
molecules and their rotational energy spectra.57 It may seem remarkable that we can treat the rotational motion
of nuclei in a molecule with the mathematical machinery of the orbital angular momentum. The reason illustrates
one of problem-solving habits youre learning: how to solve problems by appropriating techniques and results
from problems youve already solved.
Figure 4.12.1 shows the simplest model of the nuclear geometry of a rotating diatomic molecule. In this rigidrotor model , we pretend that the molecule consists of two point masses at the ends of a (massless) rod of fixed
length Re .58
(a) Write down a classical expression for the total energy E of the rigid rotor in Fig. 4.12.1 in terms of the
transverse velocities of the two masses. This is the classical Hamiltonian of the system. Next write down the
TISE of this system.Hint: In the rigid-rotor model, the total energy is equal to the kinetic energy of rotation.
The TISE of part (a), which you wrote in the coordinates r1 and r2 of the two masses, is a second-order partial
differential equation in six variables (ugh). Thankfully, we can reduce it to an equation in three variables: the
coordinates of relative motion of a single fictitious pseudo-particle, by separating out the center-of-mass motion.
56 Commentary: Dont be satisfied with just any old graph! Explore alternative graphics commands until you find one that
vividly illustrates your conclusion about these symmetry properties. Really think about what the graph is telling you about
the potential energy and its stationary states.
57 Jargon: A diatomic molecule is one with two nuclei. In a heteronculear molecule, the two nuclei are different; in a
homonuclear molecule, the nuclei are identical.
58 Details: The separation R between the two nuclear centers in a diatomic molecule is called the internuclear separation.
e
In general, the nuclei of a diatomic molecule undergo translational motion (movement of the whole molecule through space)
and two kinds of internal motion: vibrations along the internuclear axis and rotations about an axis that is perpendicular
to the internuclear axis and passes through the center of mass of the nuclei. (Chap. 9 in Understanding Quantum Physics
shows how to approximate the vibrations of the nuclei by a 1D simple harmonic motion.) The forces on the nuclei are such
that (to a very good approximation) the rotational and vibrational motions can be studied independently. Here we want to
study rotations, so well ignore the translational and vibrational motion by assuming that the internuclear separation is fixed
at its equilibrium value, Re . This assumption is the rigid rotor approximation or, more colorfully, the dumbbell model of a
molecule.

JQPMaster

Version: 8.35

Printed: August 17, 2010

324

4.11. Exercises & problems for Chap. 4

Figure 4.12.1. A crude (but effective) model of the


nuclei of a diatomic molecule. The bound states
of the molecular electrons are spatially localized in the
vicinity of the nuclei. The molecule rotates about an
axis perpendicular to the line separating the two nuclei
(the large blobs in the figure). The separation between
the nuclei is fixed at Re .

(b) Transform your expression for the energy from part (a) to center-of-mass coordinates, writing your result in
terms of the total angular momentum of the rotor, L; the constant separation between the nuclei Re ; and the
reduced mass of the rotor, . Use your result to determine the Hamiltonian of the relative motion of the
two nuclei.
(c) Determine the Hamiltonian eigenfunctions (, ) and bound-state energies of stationary states of the rigid
rotor. Label your eigenfunctions and energies with two quantum numbers: j for the rotational angular
momentum, and mj for its projection along ez . Hint:: You need do almost no algebra. Think carefully
about the operator in the TISE. You already know the solutions to its eigenvalue equation.
(d) Derive an expression for the separation between adjacent rotational energy levels E = Ej+1 Ej . How does
E change with increasing j? What is the degree of degeneracy of these energies? Why?
To illustrate the usefulness of your results, well consider an example from molecular spectroscopy . In a spectroscopic measurement we examine photons absorbed by (or emitted from) molecules when they undergo transitions
between stationary states (Chap. 18). Measured rotational spectra consist of lines whose wavelengths correspond
to transitions between rotational energy levels; these rotational lines are in the microwave or far-infrared range of
the electromagnetic spectrum. Well denote by the subscripts i and f the initial and final energy levels of the system.
From study of molecular symmetry, one learns that observed transitions correspond to a change in the rotational
quantum number j of one unit. This finding is called a selection rule: j = jf ji = 1.
(e) Derive an expression for the frequency (in cycles per second) of light absorbed by a diatomic molecule in the
rigid-rotor model.
(f) Figure 4.12.2 shows spectra from an experiment in which continuous infrared radiation passed through a gas
of HCl molecules. This figure shows absorption at the following wavelengths (in microns): 479, 243, 162, 121,
96. Assuming that each wavelength corresponds to a transition allowed by the selection rule j = jf ji = 1,
assign rotational quantum numbers j to each line. Verify that the corresponding frequencies are proportional
to j + 1, the quantum number of the final rotational state.
(g) Evaluate the equilibrium internuclear separation Re for HCl (in the rigid-rotor approximation). Compare
.
your answer to the experimental value, Re = 1.2744A
P 4.13.

*** Precession in quantum physics


In classical physics a charged particle with non-zero angular momentum L has an orbital magnetic moment L
that is proportional to L:
(4.13.1a)
L = L,
orbital magnetic moment.
where is a constant called the gyromagnetic ratio. If we expose the particle to an external magnetic field B, the
interaction of the magnetic moment with the field gives rise to an orientational potential energy
V B = L B,

JQPMaster

orientational potential energy.

Version: 8.35

(4.13.1b)

Printed: August 17, 2010

325

Figure 4.12.2.
A
rotational
absorption spectrum for HCl.
[From D. Bloor et
al. Proceedings of
the Royal Society
of London A 260,
510 (1961).]

Consequently the particle experiences a torque B = L B that tries to align L along the field axis. Therefore L
changes with time, the moment precessing about the field axis, rather like a little charged gyroscope. The angular
frequency of this precession is (` /L)B.
But what happens in quantum mechanics? The orbital angular momentum operator b
L does not have a well-defined
orientation . . . so what is it that precesses? To find out, well here investigate the time variation of the expectation
value of the quantum-mechanical orbital angular momentum.
Consider an electron with a spherically symmetric potential energy. Well denote its Hamiltonian in the absence
b (0) . The Hamiltonian in the presence of the field is
of the field by H
b =H
b (0) b
H
L B.

(4.13.2)

The quantity we want to look at, the mean value hLi, is a vector (Chap. 3).
(a) Derive an equation of motion for the time-rate of change of hLi for an arbitrary state (r, t).
b Does b
b What do yours answer imply about time rate of change
(b) Does b
L commute with H?
L 2 commute with H?
of hLi?
(c) For convenience, lets pick ez along axis of the external magnetic field. In terms of the angular frequency
B , write down equations for the rate of change of the components of hLi. Show that your equations
predict clockwise precession of hLi about the z axis.
(d) Does the magnitude of hLi change with time? Prove your answer. Hint: Consider the rate of change of
|hLi|2 = hLi )2 .
P 4.14.

**** Deuteronomy I: the spherical square well.


One of the most fascinating systems in nuclear physics is the deuteron (see Question 4.1 and the Aside on p. 341).
Since the 1930s, physicists knew that of all possible binary combinations of two nucleons (neutrons and protons) only
one is stable: the neutron-proton pair (np) called the deuteron. Experiments established that the force between
these particles is directed along the line between them, but no one could derive an explicit form for the potential
energy of the deuteron.59
(a) Write down the TISE for a deuteron in terms of the positions of the neutron and proton, rp and rn . Reduce
your equation to an ordinary radial differential equation in a single variable, the separation between these
particles, r |rp re |. Your equation describes a pseudo-particle of mass , the reduced mass of the
deuteron. Justify each step of your analysis. What can your deduce at this point about the stationary states
of the deuteron.
59 Commentary: Physicists interested in atoms and molecules have one advantage over nuclear physicists: they know the
analytic form of the dominant potential energy, the Coulomb potential energy Ze02 /r. In the early days of nuclear physics
much creative effort was invested in devising an appropriate potential energy for nucleons. This binding is now known to occur
via a mediating particle, the neutral pion.

JQPMaster

Version: 8.35

Printed: August 17, 2010

326

4.11. Exercises & problems for Chap. 4

In early days, experimental evidence indicated only that the forces between the neutron and the proton were central
and attractive, and that a bound state existed with the tiny binding energy |E1 | = 2.225 0.003 MeV. Being quite
inventive, nuclear physicists devised a host of more-or-less successful analytic models of the deuteron based on this
data. One of the simplest is a spherical square well of width a, the range of the nuclear force, and V0 > 0, the
well depth:
(
V0 0 < r a,
V (r) =
spherical square well.
(4.14.1)
0
r > a,
(b) The ground state of a particle with this V (r) must be spherically symmetric. (Why?) So we can set ` = 0 and
consider only s states. From the radial equation for Eq. (4.14.1) derive a transcendental equation whose roots
we can use to calculate the stationary-state energies of the deuteron. Hint: The long tedious way to do this
problem is to dive headfirst into algebra. The short way is to brainstorm: how does this problem resemble a
familiar problem youve already solved? Precisely how does it differ? How can you use part of the solution of
that familiar problem towith full mathematical justificationwrite down the answer to the present problem.
(at this stage we face a purely mathematical problem: to solve a second-order ordinary differential equation
subject to the boundary conditions for bound states. The solutions of such an equation, however you get your
hands on them, are unique.)
(c) Solve your equation from part (b) on a computer and explore quantitatively the relationship between the
well-depth and the range of the nuclear force. Using the binding energy given above, tabulate and graph the
dependence
of V0 on a.60 You will find that the number of ` = 0 bound states is determined by the constant
p
W (2/~2 ) V0 a2 .
(d) Does at least one bound s state exist regardless of the values of V0 and a? If not, state the conditions these
parameters must satisfy for the deuteron to support at least one bound state. What is the condition for two,
three, or more bound states?
Nuclei are small critters, and the range of the nuclear force is roughly delimited by their size. (The characteristic
size of a nucleus is 1 fermi = 1 fm = 1 1015 m.) Thats why we refer to the nuclear force, unlike the Coulomb
and gravitational forces, as short range. If our square-well model is to produce a force strong enough to support at
least one bound state, its well must be pretty deep. So its probably safe to assume that the depth V0 is considerably
larger than the measured binding energy |E1 |. Lets use this assumption to relate the fundamental parameters in this
model.
(e) Assuming V0 |E1 |, derive an approximate relationship between the well depth and the range of the nuclear
force. Now put a new curve on the graph you prepared in part (d) that shows the range of V0 for which your
approximate relation is valid. Hint: Heres part of the answer: V0 a2 . Turn this proportionality into an
(approximate) equality.
(f) Physicists also knew from measurements that the extent of the nuclear force is roughly a = 1.4 fm. Use this
additional information to evaluate (in MeV) the depth V0 of the spherical well (4.14.1) that conforms to the
known properties of the ground-state of the deuteron. What relationship do you find between your value for
V0 and the ground-state binding energy? How many bound s states does this well support?
As you discovered in part (c), this well is so narrow that it supports only one state with ` = 0. But that doesnt
necessarily mean it does not support a state with ` > 0. To find out whether such a state exists requires a little more
algebra than the s-state case, but the payoff is fascinating.
(g) Write down the reduced radial equation for ` > 0 in the classically allowed (CA) and classically forbidden (CF)
regions for Eq. (4.14.1) and show that its general solution is a linear combination of a spherical Bessel
function and a spherical Neumann function. Impose bound-state boundary conditions but dont normalize
your radial functions (doing so is a pain you neednt bother with if you want only the energies).
(h) Now enforce continuity of the radial function and its first derivative to derive the following rather forbidding
looking but harmless transcendental eigenvalue equation

j`1 (ka)
h`1 (i a)

=
i
,
(4.14.2a)
j` (ka)
k
h` (i a)
where h` (i a) is a hankel function, and the wave number k (in the classically allowed region) and the decay
constant (in the CR region) are
60 Read

JQPMaster

on: If youre a little rusty on how to solve such equations, see 8.8 of my book Understanding Quantum Physics.
Version: 8.35

Printed: August 17, 2010

327
r
k

2
|E|,
~2

r
and

2
(V0 |E|),
~2

(4.14.2b)

with the reduced mass of the deuteron.


Hint: Rather than enforce two continuity conditions, one on uE,` (r) and one on its derivative, just enforce
continuity of the logarithmic derivative of the radial function. Feel free to use the following handy recursion
relation for spherical Bessel functions:
d
`+1
j` () = j`1 ()
j` ().
d

(i) Derive the following eigenvalue equation for ` = 1 (p states),

1
1
1
1
cot ka
=
1+
ka
ka
a
a

(p states only).

(4.14.3)

(4.14.4)

Under what conditions on V0 and a does our square-well support at least one p state? Is this condition the same
as the one for at least on s state? Discuss the implications of your answer for bound states of the deuteron.
(j) If an ` > 0 bound state of the deuteron does exists, then its binding energy will be smaller than that of the
ground s state. The range a of the nuclear force is about 1 fm, which means that a < 1. Use this fact to show
that Eq. (4.14.4) reduces to the much simpler approximate relationship
j`1 (ka) 0

(` > 0).

(4.14.5)

What does Eq. (4.14.5) predict for the minimum value of the well-depth V0 that could support an ` = 1
bound state? Is your answer consistent with that of part (i)? Calculate the minimum well depth V0 that
would support a p state in a well of width a = 1.4 fm. What does your answer imply about the existence of
` > 0 bound states of the deuteron?
P 4.15.

**** Deuteronomy II: the exponential potential energy


The spherical well of Problem 4.14 is suffers from an unphysical discontinuity at a. One of the most successful
alternatives is the Yukawa potential energy weve investigated previously (see the Aside on p. 341). The Yukawa
function is not, alas, amenable to simple analytic analysis of the sort in this chapter. So instead heres a simpler
potential energy with the same essential physics: the exponential potential energy
V (r) = A er/a .

exponential potential energy.

(4.15.1)

We know from Problem 4.14 that the deuteron has only one bound state, an s state. So need consider only ` = 0.
Experimental data shows that the strength parameter A 32 MeV and the range of the nuclear force is about 1.4 fm.
(a) Reduce the TISE for (4.15.1) to a differential equation for the reduced radial function un (r), where Ive
dropped the subscript ` = 0. To simplify this equation, get rid of the exponentials by implementing the
change of variables
r er/2a .
(4.15.2)
Show that the resulting equation for u() is a form of Bessels equation,

2
1 d
q2
d
2
+
+
c

un () = 0,
Bessels equation,
d 2
d
2

(4.15.3a)

with appropriate definitions for c and q. The general solution of Eq. (4.15.3a) is
un () = d1 Jq (c) + d2 Jq (c).

(4.15.3b)

(b) Impose the asymptotic boundary condition for a bound state and, using the limiting properties of Bessel
functions, simplify this general solution. Then impose the condition that un,` (r) must be regular at the origin,
and show that this condition forces quantization of energy. Discuss ways you could solve your equation to
determine the bound-state energies of the deuteron.
(c) Calculate the binding energy of the ground state of the deuteron using the above values for the strength and
range parameters. How close is your answer to the experimentally measured binding energy 2.226 0.003 MeV?

JQPMaster

Version: 8.35

Printed: August 17, 2010

328

4.11. Exercises & problems for Chap. 4


If you have implemented this calculation on a computer, adjust the range parameter A to obtain better
agreement with this value.61

By now, you ought to be wondering, Just how well do these models simulate reality? If a model produces a pretty
good energy, does that mean that their underlying assumptions are right? Lets find out. One piece of experimental
evidence weve ignored so far is the measured value of the magnetic moment of the deuteron . Nuclear magnetic
moments are usually written in units of the nuclear magneton N , which is defined in terms of the mass of the
proton and fundamental constants as N = e~/2mp . Here are the measured magnetic moments of the proton, neutron,
and the deuteron (Appendix C):
p = 2.792 847 356 N , n = 1.913 042 73 N ,

and

d = 0.857 438 2308 N .

(d) So far we have assumed that the ground state of the deuteron is spherically symmetric. What relationship would
these magnetic moments have to satisfy for this assumption to be correct? Do they satisfy that relationship?
If not, how seriously do they violate it? What do you conclude about the validity of central-force models of
the deuteron?
(e) The deuteron has a non-zero electric quadrupole moment : Q = 2.82 1027 cm2 . What does fact imply
about the position probability distribution of the bound s state of the deuteron?
P 4.16.

*** 1/2 The Yukawa potential energy in plasma physics


Under certain conditions (in regions of high temperature and low density) we can accurately model the behavior of
bound electrons in a hydrogen plasma by the Yukawa function; in this field, this model is called the screened
Coulomb potential energy (or the the Debye-H
uckel potential energy ) and is written62
V (r) =

Ze20 r/D
e
,
r

(4.16.1)

where Z is the atomic number, and D is called the Debye screening length . In this problem well use tools youve
learned in this chapter to solve the TISE qualitatively. Table 4.16.1 gives the eigenvalues (in atomic units) for two
values of D.
Table 4.16.1. Data for the screened Coulomb potential energy
in Problem 4.16.

Energy level

D = 9 a0

D = 15 a0

1s
2s
2p
3s
3p
3d

0.7951
0.08831
0.08031
0.00319
0.00025

0.8731
0.1400
0.1366
0.03871
0.02195
0.01695

(a) On a graph of V`eff (r) for this V (r) for D = 15, show the bound-state energy levels. For each energy level,
calculate the classical turning points. Now superimpose on each energy level a sketch of the corresponding un,` (r). Make your sketches as accurate as possible. You will find that for some bound states you will need
to zoom in on particular regions of the potential energy to show un,` (r); try, though, to include more than
one bound state per graph (for instance, show the 2p and 3p bound states on the same graph), to illustrate the
difference due to the increase in n.63
61 Read on: If you havent previously run into Bessel functions and you want to learn about them, seek out any undergraduate
math methods book, such as Mathematical Methods in the Physical Sciences (Second edition) by Mary Boas (New York:
Wiley, 1983) or Mathematical Methods for Physicists (Second Edition) by George Arfken (New York: Academic Press, 1970).
A considerable more definitive and detailed book loaded with data is Tables of Functions by Eugene Jahnke, and Fritz Emde
(New York: Dover, 1945).
62 Read on: In your library youll find a large literature on the study of this potential energy. This extremely useful form also
applies to highly excited bound states of non-hydrogenic atoms in a plasma. Its not analytically solvable, but you can find
numerical solutions in the following articles: C. A. Rouse Physical Review, 159, 41 (1967) and F. J. Rogers, H. C. Graboske,
Jr. and D. J. Harwood, Physical Review A, 1, 1577 (1970), a perturbation-theory treatment in G. J. Iafrate and L. B. Mendelsohn,Physical Review, 182, 244 (1969), and variational treatments in C. S. Lam and Y. P.Varshni, Physical Review A, 19, 413
(1979) and K. M. Roussel and R. F. OConnell, Physical Review A, 9, 51 (1974), from which I took the numbers in this problem.
63 Suggestion: You will find it very helpful to prepare your graphs of potential energies and energy levels on a computer; then
print out the graphs and hand-sketch the radial functions.

JQPMaster

Version: 8.35

Printed: August 17, 2010

329
(b) Draw comparative sketches of the lowest bound-state un,` (r) of each angular momentum for the D = 9 a0 and
D = 15 a0 wells. Superimpose each sketch on a corresponding graph of V`eff (r) for each case.
P 4.17.

**** The Virial Theorem for any single-particle system


The Virial Theorem of 4.2.4 is a powerful tool for assessing wave functions obtained by solving the Schr
odinger
equation with model potential energies. In this problem, you can discover how to derive a form of this powerful
b + V (r). First you need to make a couple of tools.
b =T
theorem that holds for a single-particle Hamiltonian H
(a) Derive the commutation relation
(b) Show that for any stationary state,

b br V ).
b = i ~ (2T
br b
p,H

(4.17.1)

b i=0
h br b
p,H

(4.17.2)

(c) Use these results to show that for any stationary state,
hT i =

1
hbr V i
2

(4.17.3)

(d) Show that for atomic hydrogen in the pure-Coulomb potential energy, Eq. (4.17.3) reduces to
1
hT in,` = hV in,` .
2

JQPMaster

Version: 8.35

(4.17.4)

Printed: August 17, 2010

330

Complement 4.A. Additional examples

Complement 4.A. Additional examples


Facts are one thing, he said slowly. Interpretation is something else.
Putting facts in the right framework.
Going after Cacciato by Tim OBrien

Example 4.12 (How to exterminate unwanted first derivatives.)


This example illustrates a general technique for eliminating unwanted derivative terms from a differential
equation by deriving the reduced radial equation, Eq. (4.4.4), p. 282. The term I want to get rid of is the
first derivative in the radial equation, (2/r) dR(r)/dr, which appeared when I applied the radial kinetic-energy
operator to the radial function Rn,` (r).64
Heres the plan: Ill write the radial function Rn,` (r) as the product of two new functionsan auxiliary
function f (r) and the reduced radial function u(r)that is, R(r) = f (r) u(r). I do this in the hopes that
I can find a form for f (r) such that the differential equation for u(r) has no first-derivative term.
To find f (r), I must insert R(r) = f (r)u(r) into the radial equation (4.4.1a), isolate the terms that contain
du/dr, and set whatever multiplies du/dr equal to zero. Only the radial kinetic-energy operator (4.4.1b) can
generate such terms, so Ill start with it. The prefactor ~2 /2m will just get in the way, so for the moment Ill
omit it by writing Eq. (4.4.1b) as

2m b rad
d d
d2
2 d
T (r) = 2 +
=
+
2
~
dr
r dr
dr
| {zdr}
first term

2 d
.
dr }
|r {z

(4.A.1)

second term

Ill take the derivative terms in Eq. (4.A.1) one at a time, starting with the first term:
h

d
d
d
df
du
f (r) u(r) =
u(r)
+ f (r)
dr dr
dr
dr
dr

(4.A.2a)

= u(r)

d2 f
du df
d2 u
df du
+
+ f (r) 2 +
dr2
dr dr
dr
dr dr

(4.A.2b)

= u(r)

d2 f
d2 u
du df
+ f (r) 2 + 2
.
2
dr
dr
dr dr

(4.A.2c)

Now I go to work on the second term in Eq. (4.A.1):

2 d
2
du
df
f (r) u(r) =
f (r)
+ u(r)
.
r dr
r
dr
dr

(4.A.2d)

b rad :
Adding Eq. (4.A.2c) to Eq. (4.A.2d), I get the net effect of the derivatives in T

2 d

d
d2 u
d2 f
df
du df
2
f
(r)
u(r)
+
f
(r)
u(r)
=
f
(r)
+
u(r)
+
u(r)
+
2
+
f
(r)
,
dr2
r dr
dr2
dr2
dr
dr
dr
r
{z
}
|

(4.A.3)

where the underbrace emphasizes the stuff that multiplies the first derivative du/dr. Remember: the job
of f (r) is to eliminate this term from my final equation for u(r). So I must require that f (r) satisfies
2

df
2
df
1
+ f (r) = 0 =
+ f (r) = 0.
dr
r
dr
r

(4.A.4a)

The form of this first-order differential equation is standard: its solution is f (r) = c/r, where c is an as-yetundetermined constant of integration. [We need not worry about the value of c, since it will cancel out of
our final equation for u(r).] So the relationship between the radial function R(r) and the reduced radial
function u(r) is R(r) = c u(r)/r.
b rad act on
Now I can derive the differential equation for the reduced radial function u(r). Letting T
R(r) = c u(r)/r, I get
64 Notation: Since this example contains some moderately complicated algebra, Ive simplified the notation by temporarily
dropping all subscripts on radial functions and energies.

JQPMaster

Version: 8.35

Printed: August 17, 2010

331
i
~ 2 1 d2 u
u(r) = c
.
r
2m r dr2
The other terms in the radial equation Eq. (4.4.1a) acting on R(r) = c u(r)/r give

2
i
~ `(` + 1)
c
1 ~2 `(` + 1)
+
V
(r)

E
u(r)
=
c
+
V
(r)

E
u(r).
2mr2
r
r
2mr2
b rad
T

hc

(4.A.5a)

(4.A.5b)

Although Eq. (4.A.5b) looks a bit cluttered, it poses no difficulties, because every term in braces on the left-hand
side is multiplicative: none of these terms introduce derivatives. Setting the sum of these expressions to zero,
as prescribed by the radial equation (4.4.1a), and dividing out the common factor c/r, I get the long-sought
equation for the reduced radial function u(r) [Eq. (4.4.4), p. 282]

~2 `(` + 1)
~ 2 d2

(4.A.6)
+
+
V
(r)

E
un,` (r) = 0
n,`
2m dr2
2mr2
where Ive restored the subscripts that label a particular solution.
As desired, choosing the factor f (r) as the solution of Eq. (4.A.4a) gives an equation for u(r) that contains
a second derivative but no first derivative. Since the integration constant c is arbitrary, we might as well choose
c = 1. With this choice, the definition of and equation for the reduced radial function are Eqs. (4.4.3), p. 282.

Example 4.13 (Calculating the mean energy of a SAME from the reduced radial function.)
Some approximate and numerical methods for solving the reduced radial equation yield an expression (or table
of numbers) for un,` (r) but not for En,` . Nevertheless, we can always calculate En,` as the expectation value
of the Hamiltonian with respect to our approximate wave function. (This tactic is related to the variational
method of Chap. 15.) In Complement 4.D, Ill evaluate the energy in this way for the 2s valence electron in
the ground-state of a lithium atom (Li: Z = 3). Since that state has ` = 0, the algebra required to get the
energy is so especially simple. Here Ill show you how to cope with ` > 0.
I want to evaluate the expectation value of the energy for a state represented by n,`,m` (r):
b rad +
b | n,`,m i = hn,`,m | T
En,` = hn,`,m` | H
`
`

1 b2
L + V | n,`,m` i.
2mr2

(4.A.7)

This is a three-dimensional integral over the spherical coordinates (r, , ).

Tip. The key to successful evaluation of such expectation values is to be be organized and systematic.
Consider each term one at a time, starting with the term that is algebraically the simplest.
The simplest term in Eq. (4.A.7) is the one that contains V (r). So Ill evaluate it first. Using the definition
of SAMEs and orthonormality of spherical harmonics, I get
Z
hn,`,m` | V | n,`,m` i =
u2n,` (r) V (r) dr.
(4.A.8)
0

The next simplest term is the one that contains the angular kinetic energy operator . To evaluate this term,
Ill use the eigenvalue equation for b
L 2:

n,`,m`

1
~2 `(` + 1)
1 b 2

=
R
R
,
n,` 2 n,`
n,`,m`
2mr2
2m
r
r
~2 `(` + 1)
=
2m

u2n,` (r)

1
dr,
r2

eigenvalue equation for b


L 2.

(4.A.9a)

definition of un,` (r).

(4.A.9b)

Finally Ill tackle the term that contains radial kinetic-energy operator . Here, again, I can use orthonormality of the spherical harmonics to evaluate the angular integrals:

E Z
D
b rad
b rad (r)Rn,` (r) r2 dr
Rn,` (r)T
(4.A.10a)
n,`,m` T
n,`,m` =
0

JQPMaster

~2
2m

Version: 8.35

un,` (r)
0

d2 un,` (r)
dr.
dr2

(4.A.10b)

Printed: August 17, 2010

332

Complement 4.A. Additional examples


Adding my results for each term in Eq. (4.A.7), I get a sum of radial integrals which I can evaluate, analytically
or numerically, for any V (r) and un,` (r):
Z
Z
Z
d2 un,` (r)
~2 `(` + 1) 2
~2
1
(4.A.11)
hEin,` =
un,` (r)
dr
+
u
(r)
dr
+
u2n,` (r) V (r) dr
n,`
2m 0
dr2
2m
r2
0
0

Example 4.14 (A simpler equation for the radial dependence.)


To illustrate how generic properties can simplify solution of equations, Ill here use what we learned in 4.6.3
about the asymptotic behavior of u(r) to find an equation for the unknown r-dependence of this function.
First, I define an auxiliary function g(r) by
u(r) = g(r) er

g(r) u(r) e+r ,

(4.A.12)

where g(r) is necessarily a finite polynomial. (Why?) To derive an equation for g(r), I plug Eq. (4.A.12) into
the reduced radial equation [Eq. (4.6.6), p. 290] and perform some modest algebra:
2

~ `(` + 1)
~2 d2 g
~2 dg

+
(4.A.13)

+
+
V
(r)
g(r) = 0.
2m dr2
m dr
2mr2
You may be wondering, What earthly good is Eq. (4.A.13)? After we went to all the trouble to get rid of the
first derivative term in the radial equation, were now confronted by an equation that contains a first derivative
term. This is progress? Actually, it is. Because the definition of g(r), Eq. (4.A.12), removes the decaying
exponential e+r from u(r), the differential equation (4.A.13), unlike the reduced radial equation itself, can be
solved directly by the method of power series.
J

Example 4.15 (Buckeyballs made simple.)


One of the most remarkable molecules in nature consists of a spherically symmetric cluster of 60 carbon atoms
(!). Named after its discover R. Buckminster Fuller, this molecule is called a fullerene or buckeyball
[see Fig. 4.1, p. 290]. First discovered in the early 1980s, buckeyballs are known to exist in naturein, among
other places, micrometeorite craters on satellites. Remarkably, properties of the SAMEs of a buckeyball can
be accurately determined from the model potential energy
V (r) = V0 e

(rr0 )2

(4.A.14)

where V0 , , and r0 are parameters. Matching this model to experimental data gives r0 = 7 a0 , = 1/2,
and V0 = 4 Eh , values I used to plot the C60 potential energy in Fig. 4.1, p. 290.65
J

Example 4.16 (Dense plasmas made simple.)


A dense plasma consists of a large number of highly stripped ions and free electrons, all in a confined region.
An excellent model of the environment of one of these electrons is the Yukawa potential energy [see Complement 11.5],
er
,
Yukawa potential energy
V Y (r) = A
(4.A.15)
r
which in plasma physics is called the static screened Coulomb potential energy . In applications to a particular plasma, one chooses the parameters A > 0 and > 0 to reproduce measured properties of the plasma.J

65 Commentary: A true polymath, R. Buckminster Fuller (1895-1983) was an engineer, inventor, designer, architect, writer,
educator, philosopher, and poet, and invented the geodesic dome which gave this strange molecule its name.

JQPMaster

Version: 8.35

Printed: August 17, 2010

333

Complement 4.B. Bound- and continuum-state radial functions


Beauty is a kind of harmony and concord of all the parts to form a whole which is
constructed according to a fixed number, and a certain relation and order,
as symmetry, the highest and most perfect law of nature.
Alberti

In this chapter, I remarked on the differences between bound states and continuum states. In one respect, these
types of states are profoundly different. This difference arises from the different boundary conditions for bound-state
and continuum stationary-state wave functions.
Any bound-state wave function (r, t) must be normalizable; once normalized, this function can be interpreted as
a position probability amplitude. But its not possible to normalize the wave function of a continuum stationary state,
because a particle in a continuum state isnt confined to a finite region of space for all time. As you may recall from
1D quantum mechanics, the inability to normalize a continuum-state wave function significantly affects its interpretation. For 3D SAMEs, our major concerns are normalization and interpretation of the radial functions RE,` (r). This
complement discusses these points.
4.B.1 Bound versus continuum states
Definition (bound state) In a bound state the particle is confined (spatially localized) in the vicinity of the
potential energy, which must be attractive. At any time, the probability of finding the particle far from the region of
where the potential energy is appreciable is vanishingly small.
Definition (continuum state) In a continuum state the particle is not spatially localized. As its wave function
evolves, the particle might be found anywhere. Continuum states are scattering states, 3D counterparts of the
1D states used to calculate transmission and reflection coefficients.66

The probability density of a stationary state doesnt depend on time. For such a state, the distinction between
bound- and continuum-state wave functions reduces to a distinction between Hamiltonian eigenfunctions. This brings
us to the difference between normalization conditions mentioned above:
Rule: The Hamiltonian eigenfunction of a bound-state must be normalized; the Hamiltonian eigenfunction of a continuum
state cant be normalized.
This difference means that the stationary-state energies of bound states are quantized (discrete), while the energies
of continuum states are not. Figure 4.1, p. 267 illustrates this difference for a generic central potential energy that
supports two bound-state energies.67
In this figure, I set the zero of energy at the top of the potential energy well, so V (r) 0 as r . With this
choice, the bound-state energy spectrum consists of discrete negative numbers, and the continuous energy spectrum
consists of all positive numbers E 0. The different nature of bound- and continuum-state energies significantly
alter our goals in problem solving. For a bound state, we dont know the energies: to obtain them, we must solve the
TISE. But for a continuum state were free to choose any energy E > 0 we want. Our reason for solving the TISE
for a continuum state is to get the eigenfunction E (r) for a known energy.
The most common use of a continuum state is to approximate a beam of particles in a scattering experiment.
The value of E is the (approximately sharp) energy of the particles in the beam. Solution of the TISE yields the
asymptotic behavior (r ) of the corresponding eigenfunction E (r), from which we can calculate the cross
section. This asymptotic behavior is specified by boundary conditions we impose on the solution.68
66 Read on: You can find a very detailed discussion of continuum states and 1D scattering theory in 1D in 8.3 of Morrison
(1990).
67 Read on: In 1D, any attractive potential energy supports at least one bound state. This is not true in three dimensions
(see Complement 4.C). For example, a spherically symmetric (radial) well of depth V0 and radius r0 , defined as V (r) = V0
for 0 r r0 and V (r) = 0 for r r0 , supports no bound states unless V0 > ~2 2 /(8mr02 ). For more on this subject, see
Potentials and Bound States, by Walter F. Buell and B. A. Shadwick, American Journal of Physics 63, 256, (1995).
68 Details: For a continuum state,
E,`,m` (r) corresponds to scattering of a particle with kinetic energy E by V (r). The
energy of particles in the incident beam is set in the laboratory (by the device that emits them). The key physical quantity
is the probability that the incident particle will be scattered by the target and emerge in a direction (0 , 0 ). We can calculate
this probability, the differential cross section, from a Hamiltonian eigenfunction by imposing explicit asymptotic boundary
conditions on RE,` (r) as r that conform to the prescription of Eq. (4.B.1b).

JQPMaster

Version: 8.35

Printed: August 17, 2010

334

4.B.2 Boundary conditions on bound- and continuum-state radial functions

4.B.2 Boundary conditions on bound- and continuum-state radial functions


For a bound-state SAME of a rotationally invariant 3D system, the boundary conditions on the radial function RE,` (r)
are [Eqs. (4.2.8), p. 270]: 69
Rn,` (0) = a finite number
Rn,` (r) 0.

boundary conditions for


a bound state (3D)

(4.B.1a)

Boundary conditions on bound-state Hamiltonian eigenfunctions force energy quantization. For a SAME, the quantized energies En,` emerge from solution of the radial equation, so the boundary conditions that force their quantization are Eqs. (4.B.1a).70
For a continuum-state SAME, we require the radial function RE,` (r) to be bounded everywhere, including at the
origin:
boundary conditions for
|RE,` (r)| = a finite number for all r
(4.B.1b)
a continuum state (3D)
Key point! Whether the state is bound or not, the radial function RE,` (r) must not blow up as r .
I Aside. Limiting behavior of continuum radial functions.
In 4.6.4 we discussed the behavior of un,` (r) near the origin: uE,` (r) r`+1 . This behavior is common to
bound and continuum states. Where bound and continuum states differ is in the asymptotic limit. According
to Eq. (4.4.6), p. 282, as r a bound-state function un,` (r) must approach zero. By contrast, the most
we can require of a continuum-state function uE,` (r) in this limit is that it be bounded
(finite) for all r.

For E > 0, we can write the asymptotic equation in terms of the wave number k 2mE/~ as [for contrast
see Eq. (4.6.8), p. 291]:71
d2 u
+ k2 u(r) = 0,
dr2

(for a continuum state in the asymptotic limit.)

(4.B.2a)

This is the equation of free particle.


Rule: The asymptotic form of a continuum-state SAME is that of a free particle with a sharp energy and orbital
angular momentum.
The second-order equation (4.B.2a) has two linearly independent solutions: ei k r and e+i k r . So its
general solution is the arbitrary linear combination
uE,` (r) = A ei k r + B e+i k r

general reduced radial function for a


continuum state in the asymptotic limit.

(4.B.2b)

Unlike er , both ei k r and e+i k r are bounded, we need not set either A or B to zero.
Equation (4.B.2b) is valid only for sufficiently large r that the effective potential energy V`eff (r)
[Eq. (4.4.9), p. 284 of (4.5)] is negligible compared to other terms in the reduced radial equation [Eq. (4.4.4),
p. 282]. By restricting slightly the class of potential energies we consider, we can get a general solution thats
valid at radii where the physical potential energy but not the barrier term ~2 `(` + 1)/2mr2 is negligible. Provided V (r) r(3+) as r and V (r) r(2+) as r 0 for some > 0. we can solve for uE,` (r) at r
where V (r) is negligible but `(` + 1)/r2 is not. The appropriate equation is
2

`(` + 1)
d
2

+
k
uE,` (r) = 0,
[for r where V (r) is negligible.]
(4.B.3a)
dr2
r2
The (unnormalized) solutions of this equation are special functions called the Riccati-Bessel function `
and Riccati-Neumann function n
` of order `. The general large-r form of uE,` (r) at these radii is the
linear combination
69 Details:

More precisely, as r , the radial function must go to zero faster than r2 goes to infinity.
While there exist mathematically legitimate solutions of the radial equation for energies other than the special
values En,` , the radial functions for these energies blow up at r = 0, in the limit r , or both. Eigenfunctions that blow up
arent normalizable and so are of no use to us.
71 Notation: For a continuum state, I cant attach a principal quantum number n to k. Strictly, though, k should carry ` as
a subscript.
70 Commentary:

JQPMaster

Version: 8.35

Printed: August 17, 2010

335
uE,` (r) = A ` (kr) + B n
` (kr).

[for r where V (r) is negligible.]

(4.B.3b)

This is the starting point for partial-wave scattering theory .

Complement 4.C. The number and energy-ordering of bound states


A line from Schiller came to him:
Natural necessity has entered into no compact with man.
The world is under no obligation to make sense to us.
Panama by Eric Zencey

4.C.1 The number of bound states in a 3D attractive potential energy

Rule: Any attractive potential energy in 1D or 2D supports at least one bound state. This does not hold in 3D: there
are attractive potential energies [V (r) 0 for all r] that support no bound states.72
Physicists and mathematicians have long sought general conditions that tell us whether a given potential energy V (r)
supports any bound states and, if so, how many bound states it supports. In this complement, Ill summarize some
of their major findings.73
(1) A potential energy that is bounded as V (r) < C/r for some < 2 supports an infinite number of bound
states. This case includes the pure-Coulomb potential energy Ze02 /r of the one-electron atom ( = 1)
(Chap. 5).
(2) A potential energy that satisfies V (r) > C/r2 supports
a finite number of bound states if
an infinite number of bound states if

~2
,
2m

1 ~2
C>
.
4 2m
C<

1
4

(3) For an attractive spherically symmetric potential energy V (r) = V (r) that satisfies
Z
Z
|V (r)| r dr < and
|V (r)| r2 dr < ,

(4.C.1a)

(4.C.2)

there always exists some maximum quantum number `max such that for all ` > `max the potential energy
supports no bound states. That `max is the solution of
Z
2m
2`max + 1 2
|V (r)| dr.
(4.C.3)
~
0
Equation (4.C.3) is an inequality because the left-hand side must be an integer so `max will be an integer. The
number of bound states in any attractive potential energy that obeys Eqs. (4.C.2) is finite.74
(4) For ` < `max , the number of bound states Nn,` of angular momentum ` supported by a potential energy that
satisfies Eqs. (4.C.2) is bounded by75

Z
2 2m
Nn,`
|V (r)|1/2 dr.
(4.C.4)
~
0
72 Details: A potential energy need not be attractive (negative) everywhere to support one or more bound states. Rather,
it must be attractive
R on average. In 1D, for example, any V (x) that goes to zero asymptotically and whose integral over all
space is negative,
V (x) < 0, supports at least one bound state. This condition allows V (x) to be repulsive (positive) over
some finite range(s) of x. The same is true in 2D. But in 3D the situation is different: to support one or more bound states a
potential energy V (r) must be attractive on average and must also satisfy the conditions discussed in this Complement.
73 Read on: You can find a good summary of this topic in 6.1 of Galindo and Pascual (1991). Many of these results were
obtained using the variational method, so you can find more information and recommendations for outside reading in Chap. 15.
74 Details: This conclusion does not apply to the pure-Coulomb case V (r) 1/r 2 , which violates Eqs. (4.C.2). Well deal
with this in Chap. 5.
75 Read on: This result follows from basic quantum mechanics: see Chap. 23 of Calogero (1967).

JQPMaster

Version: 8.35

Printed: August 17, 2010

336

4.C.2 The order of bound-state energies of a rotationally invariant system


Like Eq. (4.C.3) for `max , Eq. (4.C.4) is an inequality because the left-hand side must be an integer.

(5) A monotonic, attractive, spherically symmetric potential cannot support any bound states unless
Z
~
[V (r)]1/2 dr >
.
2 2m
0

(4.C.5)

4.C.2 The order of bound-state energies of a rotationally invariant system

Solving the Schr


odinger equation for a rotationally invariant system is so much easier than for a system without this
symmetry that physicists often model a potential energy thats not spherically symmetric by one that is (see Chap. 11).
Hence the importance of generic properties of the bound-state energies of a central potential (4.6). In addition to
the number of bound states such a potential can support, we would also like to know the order of bound-state
energies En,` for ` 0, 1, . . . and n = 1, 2, . . . , ` 1.76
Rule: The energy ordering of En,` depends on the number of nodes in the radial function, nr = n ` 1.
I now need to introduce yet another index (sorry): n
nr + 1 = n `. When you use the following results, remember
that, choosing the zero of energy as in Fig. 4.1, p. 267, bound state energies are negative: En,` < 0:
(1) The degree of degeneracy of each bound SAME is g(En ,` ) 2` + 1.
(2) For two bound SAMEs with the same ` and with n0 > n, the energies are ordered as En0 ,` > En,` .
(3) For any two SAMEs with `0 6= ` whose functions un ,`0 (r) and un ,`0 (r) have the same number of nodes nr , the
energies are ordered as En ,` < En ,`0 . Hence
(a) The ground-state of any rotationally invariant system is an s state (` = 0) and therefore is spherically
symmetric.
(b) The number of bound-states with quantum number ` + 1 supported by a spherically symmetric potential
energy must be less than or equal to the number of bound-states with `.
We can combine these results into the handy inequality
En 1,` En ,` En ,`+1

(for any rotationally invariant system).

(4.C.6a)

Whats so powerful about these results is that they require only that the potential energy be spherically symmetric
and strong enough to support one or more bound states.
(4) The most difficult energies to order are En 0 ,`0 and En ,` such that n
0 + `0 = n
+ `. In this case, it turns out that
if the potential is everywhere attractive, then
En ,` < En 1,`+1 ,

for all n
and ` [if V (r) < 0 for all r]

(4.C.6b)

For almost all systems of physical interest, solving the reduced radial equation requires extensive mathematical
approximations and/or numerical treatment on a computer. The more you know in advance about the results you
should get, the more likely you are to spot errors that inevitably plague such work. Were quite fortunate to have
results like those in this Complement (and in Secs. 4.6 and 4.7). Although nature refuses to give us the answers in
the back of the book, nature sometimes provides ways to determine if our answers are wrong.
76 Read on: This topic underpins the famous Aufbau principle that Neils Bohr developed in 1913 to explain the periodic
table (Chap. 12).

JQPMaster

Version: 8.35

Printed: August 17, 2010

337

Complement 4.D. The ground state of the lithium atom: a tale of three models
Is not this, perhaps, the secret of every true and great mystery,
that it is simple?
Introduction to a Science of Mythology, by C. Kerenyi

This complement illustrates the ideas and machinery weve developed for 3D quantum mechanics without getting
bogged down in the details of how to solve the radial equation. It also introduces the idea of studying real systems by
telling stories about them: The Principle of Modeling . As our test subject Ive chosen a lithium atom (Li: Z = 3)
in its ground state. Well focus on one electron in this many-electron atomthe so-called valence electronand
use an approximate analytic eigenfunction to look at radial functions and probability densities, calculate expectation
values and uncertainties, and so forth.
Neutral Li consists of three electrons, each of charge e (where e = 1.602 178 1019 C is the electronic charge,
also called the elementary charge), and a nucleus of charge +Ze with Z = 3. Each pair of these particles interacts
via the electrostatic Coulomb potential energy. So the total potential energy in the Hamiltonian of a Li atom is the
sum of electron-nucleus interaction terms plus electron-electron terms.
Electron-nucleus interaction terms. If, as in Fig. 4.D.1, we put the origin at the nucleus, then the Coulomb
potential energy between each electron and the nucleus depends only on r and so is spherically symmetric. Denoting
the electrons spatial variables by r1 , r2 , and r3 , we can write the electron-nucleus interaction terms as
V en =

Ze02
Ze02
Ze02

,
r1
r2
r3

electron-nucleus Coulomb interactions in Li,

(4.D.1a)

where Z = 3 for Li. To reduce clutter Ive introduced the convenient constant e0 , defined by
e02

e2
40

= 2.30708 1028 kg m3 s2 .

(4.D.1b)

Figure 4.D.1. Spatial coordinates


for the three electrons of atomic
lithium. The nucleus of atomic number Z = 3 is at the origin. Quantities
such as r12 |r1 r2 | are inter-electron
separations.
Electron-electron interaction terms. The other contribution to the Li potential energy are electron-electron
interaction terms, one for each pair of electrons. These potential energies depend on different variables than those
in Eqs. (4.D.1a)on the separations between pairs of electrons:77
V ee =

Rule:

e02
e02
e02
+
+
,
|r1 r2 |
|r2 r3 |
|r3 r1 |

electron-electron Coulomb interactions in Li.

(4.D.1c)

The electron-electron Coulomb potential energy is not spherically symmetric.

Including the kinetic-energy operators for each electron and all Coulomb interactions, and assuming that the
nucleus is a stationary point charge at the origin, the pure-Coulomb Hamiltonian for lithium is
77 Details: A proper treatment of any atom must incorporate the fact that electrons are identical particles. This fact imposes
a symmetry property on the Hamiltonian: it must be invarant under interchange of any two electrons. Since each electron has
the same mass, the kinetic-energy operator has this symmetry; since each has the same charge, so does the electron-nucleus
potential energy. Since the electron-electron potential energy Eq. (4.D.1c) contains an identical term for each possible pair of
electrons, it, too, has the required symmetry.

JQPMaster

Version: 8.35

Printed: August 17, 2010

338

4.D.2 Models of the ground state of lithium


2
2
2
b = ~ 21 ~ 22 ~ 23 + V en + V ee ,
H
2me
2me
2me

pure-Coulomb Hamiltonian of a Li atom.

(4.D.1d)

4.D.1 The principle of modeling


The operator (4.D.1d) is not the true Hamiltonian of a lithium atom: its the Hamiltonian of a fictitious neutral threeelectron. one-nucleus system. The nucleus of an actual lithium atom isnt a point particle and isnt fixed in space.
Electrons arent spinless particleseach electron has spin s = 1/2 (see Chap. 6). And there are interaction terms
involving various angular momenta of the electrons and the nucleus as well as relativistic-correction termsnone of
which appear in the Hamiltonian (4.D.1d). So what good is this operator?
b omits some of the physics of a true Li atom, it includes the terms that
Its value is two-fold. First, although H
represent the most important physicsthe Coulomb interactions are primarily responsible for the atoms properties.
b is sufficiently simple that, with additional approximations (Chap. 11), we can solve the corresponding
Second, H
b
time-independent Schr
odinger equation (TISE) and study the properties of Li, its behavior, and its spectra. So H
satisfies the criteria of a successful physical model . In devising Eq. (4.D.1d), we carried out one of the most common
stasks of a scientist: we modeled a complicated physical system. We invented a story about lithium
Telling stories to understand physical systems is hardly new. You used this strategy when, to determine how
far a softball would travel between a pitcher and a batter, you chose to neglect air resistance. You used it when, to
understand the oscillations of a mass on a spring, you chose to neglect anharmonic oscillations. (You modeled the
spring by a classical simple harmonic oscillator.) You used it in electricity and magnetism when, in considering an
electromagnetic field, you neglected the small magnetic component and retained only the electric component. You
used it in 1D quantum mechanics when you modeled the vibrations of a diatomic molecule as a simple harmonic
oscillator. In these and many other cases, youconsciously or intuitivelyneglected aspects of a systems physics
that you think (hope) are irrelevant.78
In science, a model is both a story and a hypothesis. Its a speculation about whats vital to a system versus
whats negligible. We test our hypothesis by calculating properties and behavior of the system based on our model
and comparing what we get to what nature tells us, to information obtained via experiments. If necessary, we revise
our model, perhaps by restoring some of what we neglected or perhaps by creating an entirely new model, and try
again. This activity is a problem-solving strategy, one of the habits of highly effective problem solvers:

Tip. The Principle of Modeling. When confronted with a complicated, unfamiliar system, devise one or more
models of the system. Seek models that retain the systems most important physical features but neglect enough of
its complications that you can solve the Schr
odinger equation for the model Hamiltonian.

4.D.2 Models of the ground state of lithium


The model implicit in the Hamiltonian (4.D.1d), which Ill call the pure-Coulomb model invokes the approximation
of infinite nuclear mass: that the nucleus is a point mass fixed at the origin (Chap. 5).79 A quick search for known
properties of Li reveals several reasons why this model is sensible and promising:
The mass of the Li nucleus is about three times the mass of a proton.
The mass of a proton is about 2000 times the mass of an electron.
78 Example:

Lets return for a moment to the aforementioned softball. Surely the most important influence on the time it
takes the softball to get from the pitcher to the catcher is gravity. Our first model might be a sphere of uniform mass density
moving only under the influence of gravity. If this model isnt accurate enough, we could embellish the story. Maybe the ball
is a cheap one, in which case we might allow for deviations from spherical symmetry or variations in mass density. Maybe
theres a moderate wind, in which case wed incorporate the wind and air resistance. Maybe theres a tornado, in which case we
should go hide and forget about modeling. In every case (except the last) we seek to reduce the complicated problem (solving
Newtons second law for the real pitch and catch) to a problem we can, perhaps with further mathematical approximation,
solve (projectile motion).
79 Jargon: The phrase infinite nuclear mass does not mean literally that the mass of the nucleus is infinite; rather, this
phrase alludes to the assumption that the nucleus is fixed in space. (A particle that never moves no matter what forces act
on it would have to have an infinite mass.) This approximation works extremely well for matter; in Chap. 5 well see that the
correction to the energies of atomic hydrogen due to the finite mass of the proton are incredibly small. In Chap. 14 well explore
the corrections due to the finite nuclear size, which are negligible for light atoms.

JQPMaster

Version: 8.35

Printed: August 17, 2010

4.D.3 Models for the valence electron in lithium

339

The size of the nucleus (roughly 1015 m) is several orders of magnitude smaller than the size of the atom
itself (roughly 2
A = 2 1010 m).80
The de Broglie wavelength of the proton, = h/p = h/mv, is much smaller than that of the electron, because
the wavelength of a particle is inversely proportional to its mass.
Trouble is, even within the pure-Coulomb model we cant solve the TISE of Li exactly. Furthermore, we dont
yet know how to cope with a system of more than one electron. But wait a minute: our goal is not to understand
everything about lithium; our goal is to understand the physics of the valence electron.
Definition (valence and core electrons) The valence electron is the bound electron that, on average, is most
likely to be found farther from the nucleus than the other electrons. The other electrons are called core electrons.
In terms of the shell model of the atom (Chap. 11), a valence electron is an electron in the outermost shell,
which for Li has n = 2. Like other alkali atoms (the Aside on p. 341), the valence electron in Li has zero orbital
angular momentum (` = 0). Spectroscopists refer to an electron with ` = 0 as a s electron and say that the valence
electron in Li occupies a 2s state. So our goal is to understand the physics of the 2s electron in Li.
To achieve this goal, we must devise a model of the valence electron in Li. We would prefer a spherically symmetric
model potential energy so we can use the machinery of rotationally invariant systems from Chaps. 3 and Chap. 4.

4.D.3 Models for the valence electron in lithium


The bare-nucleus model. A really simple model for the 2s electron in Li would just ignore the other two electrons,
in which case the 2s electron would experience only the field of a bare, triply-charged Li3+ nucleus:
Definition (The bare-nucleus model.) The hypothesis of the bare-nucleus model is that the essential physics
of the valence electron in lithium arises from its interaction with the nucleus.
While appealing simple, this model makes little sense. I mean, surely the core (1s) electrons play some role. This
model approximates the physics of an electrically neutral system by an ion. Lets try something else.
The one-electron model. We must somehow include the effect of the core electrons on the valence (2s) electron.
The main thing the core electrons do for the valence electron is to shield it from the full attractive force of the
Z = 3 nucleus. Called screening (or shielding ), this reduction of the nuclear attraction plays a major role in manyelectron atoms (Part III). A cartoon of a model that takes account of screening would show a nucleus of charge +3e
surrounded by a spherical shell of charge 2e (due to the two other electrons) and, outside this shell, the valence
electron with charge e. In this cartoon, the valence electron is localized outside the nucleus and the other two
electrons, so it experiences a Coulomb potential of net charge +3e 2e = +e. Weve now concocted a second story
about lithium:81
Definition (the one-electron model.) The hypothesis of the one-electron-atom model is that the core electrons completely screen the valence electron from the nucleus, leaving it in the pure-Coulomb potential energy e02 /r.
The mathematical translation of these two stories are the model potential energies
V (r) =

eff 2
Z2s
e0
,
r

pure Coulomb potential for a 2s electron,

(4.D.2a)

eff
where the so-called effective nuclear charge seen by the 2s electron, Z2s
, is
80 Jargon: Im using the word size rather loosely. Since quantum states are represented by wave functions which correspond
to uncertain positions, an atom or a nucleus does not literally have a size. In Chaps. 9 and 11, however, Ill define a quantity
that can meaningfully, albeit approximately, be interpreted as the size of an atom in a particular bound state. A simple example
is the radius such that the integrated probability within a sphere of this radius is, say, 0.99.
81 Details: To understand this point more fully, check out Fig. 4.D.3, p. 342. Note that the position probability density
of the 2s electron is largest outside the probability density of the 1s core electrons, while the 1s density is largest very near
the nucleus at the origin. So the (core) electrons effectively screen the 2s electron from the full attraction of the nuclear
charge 3e. The 2s electron feels a net force that, although attractive, is not as strong as that of a bare Li nucleus. This idea
supports a model electron-electron interaction potential energy for the valence electron of Li that is spherically symmetric. The
core electrons that screen the 2s electron are in a closed shell, so the net repulsive force they exert on this electron is, to an
excellent approximation, radial (see Chap. 11). Such models are widely used in the study of alkali atoms, as they offer that
combination of simplicity and accuracy that we most desire in physics.

JQPMaster

Version: 8.35

Printed: August 17, 2010

340

4.D.3 Models for the valence electron in lithium


(
eff
Z2s

3,
1

in the bare-nucleus model


in the one-electron model.

(4.D.2b)

The one-electron-atom model is surely more realistic than the bare-nucleus model. But its too classical. An
electrons wave functions decay to zero as r : its unrealistic to model an electron as strictly confined to a shell
whose radii are a few angstroms. A more serious defect is the premise that the 2s valence electron is never found
near the nucleus, that it never penetrates the cloud of negative charge due to the core electrons. Were on the
right track, but we need something a little more realistic.
Both models suffer from the same weaknesses: their dependence on r too inflexible: its constrained to purely 1/r.
eff
One way to improve on these models is to allow Z2s
to vary with r, so V (r) will vary from +e02 /r near the nucleus
2
to +3e0 /r far from the nucleus, like the solid curve in Fig. 4.D.2. Thus we arrive at our third story about Li:
Definition (the screened Coulomb model.) The hypothesis of the screened Coulomb model is that the core
electrons incompletely screen (or shield) the nuclear charge in a way that depends on the radial coordinate of the
valence electron.
0

1.0

-10

0.8
exponential screening factor

Li model

potential energy

-1r
Li
-20

-3r

-30

0.4

0.2

-40

-50
0.0

0.6

0.1

0.2

0.3

0.4

0.5

0.6

0.0
0.0

0.2

0.4

0.6

0.8

1.0

Figure 4.D.2. A model potential energy and screening function for the valence electron in an alkali
atom. Left panel: The model potential energy in Eq. (4.D.4b). Also shown are pure Coulomb potentials
V (r) = Z/r for Z = 1 (upper dash curve) and Z = 3 (lower dash curve). All quantities are in atomic units:
V (r) in Hartree and r in Bohr. Right panel: The exponential screening function er for the screened Coulomb
model,(4.D.3).
We can translate this hypothesis into mathematics by multiplying the pure-Coulomb potential energy e02 /r by
some continuous function that goes to 1 as r 0 and to 0 as r . A simple function with this behavior is the
exponential er , which gives for the valence electron (as indicated by the subscript val)82
Vval (r) =

er
,
r

screened-Coulomb potential.

(4.D.3)

The real, positive screening constant controls the extent to which the core electrons screen the +3e02 /r attractive
electron-nuclear potential energy as r of the valence electron varies. Introduction of the parameter is a great way
to increase the flexibility of a model: we can now adjust to make our model as realistic as possible. The potential
energy (4.D.3) retains the highly desirable property of spherical symmetry, which lets us go directly to the radial
function for the SAMEs of the 2s electron.
In the rest of this Complement, Ill show you results calculated from a 2s radial function based on a slight variation
on Eq. (4.D.3). After playing around with various analytic forms like Eq. (4.D.3), I settled on the following function
(in atomic units: Appendix F):83

(Z 1)
Z
1
Vval (r) = +
1 (r + 2) er ,
model potential energy for a valence electron,
(4.D.4a)
r
r
2
82 Details:

This function is a variant of the widely used Yukawa potential energy discussed briefly in the Aside on p. 341.
This form, which is based on classical electrostatics, is more flexible than Eq. (4.D.3) and gives pretty good
results for neutral atoms and ions with any atomic number Z. The expression in Eq. (4.D.4) shows how it modifies the
pure-Coulomb potential energy Z/r. For plots of this model function, see Figs. 4.D.2 and 4.D.3.
83 Commentary:

JQPMaster

Version: 8.35

Printed: August 17, 2010

4.D.3 Models for the valence electron in lithium

341

where Z is the atomic number of the atom and is a parameter. For (Li: Z = 3), this function is
Vval (r) =

(r + 2) r
1

e
,
r
r

model potential energy for teh 2s valence electron in Li,

(4.D.4b)

where, for Li, Z = 3. Comparison of the calculated energies based on this potential energy to measured spectra of Li
gave best agreement for = 5.38 a1
0 . Part (a) of Fig. 4.D.2 shows that the potential (??) behaves correctly in the
limits of zero screening (r 0) and complete screening (r ).84
I Aside. An atomic physics primer.
Lithium is one of the alkali atoms. The distinguishing feature of the alkalies (which include also sodium,
potassium, rubidium, cesium, and francium) is the distribution of their electrons. Each alkali has a lone
outer electron, the valence electron, which is localized primarily at larger r than the other inner
electrons, the core electrons.
Inherent in this description is the shell model of the atom , a story beloved of high-school and freshman
chemistry teachers. In this model, we imagine assigning each electron to a subshell. By definition, a
shell is identified by principal quantum number n, while a subshell is identified by n and `. So to assign
an electron to a subshell means to associate the electron with particular values of n and `. In so doing we
must obey the Pauli exclusion principle, according to which no more than two electrons can share the
same three quantum numbers n, `, and m` (Chap. 6).
In Li, the two core electrons are assigned to the 1s subshell, so each has n = 1, ` = 0, and m` = 0.
According to the Pauli Principle, the valence electron must go into the 2s subshell; it has n = 2, ` = 0, and
m` = 0. The core electrons are localized primarily near r = 0that is, their probability density is largest
near the origin. The 2s electron is localized at larger values of r than the 1s electrons.

Aside. Yukawa and the deuteron.


If a neutron and a proton get close enough together, they can form a stable two-particle system called a
deuteron. The bound states of the deuteron (as for any two-particle system) depend only on the relative
motion of the proton and the neutron. (The motion of the deuteron as a whole is just that of a free particle.)
By transforming the Schr
odinger equation to center-of-mass coordinates (see 5.3 of Chap. 5), we can
transform this problem into that of a single pseudoparticle whose mass equals the reduced mass of the
deuteron. But an explicit functional form for the potential energy due to the nuclear force, the analog of the
Coulomb potential energy for electric force, is unknown. So one must use a model potential energy in the
Hamiltonian of the deuteron. For years nuclear physicists modeled the nuclear force by the potential energy
V (r) = A

er
,
r

Yukawa potential energy

(4.D.5)

Named after its inventor, this model is called the Yukawa potential energy . One can choose the strength
parameter A and screening constant to reproduce the experimentally determined energy of the (single)
bound state of this system, and the resulting potential energy used to calculate other physical quantities.
(For more about Yukawa and the Yukawa potential energy, see Problems 4.14 and 4.15 as well as Chap. 11.)
84 Details: To assess various models, I compared the energy and other properties of the 2s electron in Li to whatever information
I could find in the literature. For instance, I compared the bound-state energy to the value determined by accurate spectroscopic
measurements. A fantastic resource for such data is the web site maintained by the National Institute of Standards and
Technology. As of 2009, its URL is http://physics.nist.gov/PhysRefData/ASD. For other properties such as the mean radius
of the 2s electron, I compared to results from the most accurate, sophisticated computer calculations of this quantity. This
information appears in papers published in journals such as Physical Review A and The Journal of Chemical Physics.

JQPMaster

Version: 8.35

Printed: August 17, 2010

342

4.D.4 A radial function for the valence electron in Li

4.D.4 A radial function for the valence electron in Li

Solution of the radial equation for the potential energy (??) yields the 2s radial function R2s (r) and radial probability
rad
density P2s
(r) in Fig. 4.D.3 (see the Aside on p. 344).
Noteworthy features of Fig. 4.D.3:
(1) Boundary conditions. The function R2s (r) goes to zero as r and to a (large negative) constant as
r 0.
(2) The radial probability density. While R2s (r), goes rapidly to a negative constant as r 0. the radial
rad
probability density P2s
(r) goes to zero in this limit. In the other, asymptotic limit r , this function
decays rapidly to zero.85
(3) Radial nodes. The function R2s (r) has one node (nr = n ` 1 = 2 1 = 1). This node occurs at
r = 0.74 a0 , which ensures that the area under the curve R1s (r)R2s (r)r2 , the integrand in the orthogonalrad
ity integral hR2s | R1s ir equals zero. On either side of the node, P2s
(r) has maxima at r = 0.22 a0 and 3.0 a0 .
The outer peak is much larger than the inner peak. So the valence electron is localized primarily far from the
nucleus, but has a tiny probability density very near the nucleus.
1.5

0.5

R 2 s HrL
1s

0.0

Pn,{ HrL

-0.5

VHrL
0.5

-1.0

2s

-1.5

-2.0

10

10

r Ha L

r Hao L

Figure 4.D.3. The radial function and probability density for the 2s electron in Li. Left
panel: The radial function R2s (r) for the 2s valence electron in the ground state of Li. Also shown
rad
(dashed curve) is the physical potential (??). Right panel: The radial probability density P2s
(r)
for a 1s and the 2s electrons of Li in its ground state [Eq. (4.D.8), p. 344].

Example 4.17 (The shell probability for the 2s electron in Li.)


rad
As noted in 4.3, the measurable quantity thats conceptually closest to Pn,`
(r) is the shell probability , which
for the 2s electron is [Eq. (4.3.10a), p. 279]
Z
shell
P2s
(r0 )

r0 + 1
r
2
r0 1
r
2

rad
P2s
(r) dr.

(4.D.6)

shell
Figure 4.D.4 shows P2s
(r0 ) as a function of the position r0 of a shell of thickness r = 0.0001 a0 . Careful
rad
comparison to P2s (r) in Fig. 4.D.3 shows that qualitatively, the radial probability density is identical to the
J
shell probability.86

85 Details: More precisely, for the wave function in Eq. (4.D.8), p. 344 (see the Aside on p. 344), the asymptotic behavior of
the radial function is R2s (r) e0.630 r /r as r .
86 Details: In preparing Fig. 4.D.4, I wanted to simulate the infinitesimal thickness in the formal definition of P rad (r), so
n,`
I chose a very small shell thickness r, a value roughly equal to the size of the points in the figure. But even had I chosen
shell
rad
a much larger thicknesssay r = 0.01 a0 the shape of P2s (r0 ) would have been the same as that of P2s (r). So, even
rad (r) cant actually be measured, it does give meaningful information about the likelihood of finding the 2s electron
though Pn,`
at various radial distances from the nucleus.

JQPMaster

Version: 8.35

Printed: August 17, 2010

4.D.4 A radial function for the valence electron in Li

343
scaled P2 s HrL HsolidL and Pshell
2 s HrL Hopen circlesL

0.0006

shell thickness = 0.002

0.0004

0.0003

0.0002

0.0001

0.0000
0
2
4
6
8
10
0.0005

Figure 4.D.4. The probability of detecting an electron


in a spherical shell at r0 .
The thickness of the shell is
r = 0.002 a0 . The solid curve
is the radial probability density
scaled at the peak to equal the
shell probability, values of which
are given by open circles. (Note
the vertical scale compared to
that in Fig. 4.D.3.)
I

radius Ha0 L

Example 4.18 (Quantitative validation of the model radial function.)


I can evaluate the bound-state energy E2s as the expectation value
b + V | 2s i,
b | 2s i = h2s | T
E2s = h2s | H
r

(4.D.7a)

where V is the potential energy Eq. (??). Because the state is a SAME, the implied triple integral (over r, ,
and ) reduces to the radial integral (see Example 4.13 in Complement 4.A)
Z
h
i
b rad + Vval (r) R2s (r) r2 dr,
E2s =
R2s (r) T
(4.D.7b)
0

b rad is the radial kinetic-energy operator [Eq. (4.4.1b), p. 281]. Numerical evaluation gives E2s =
where T
0.19845 E h , where E h denotes Hartree (Appendix F): 1 E h = e02 /a0 = 27.2114 eV. Converting units gives
E2s = 5.40017 eV, which agrees spectacularly well with the experimentally determined energy 5.39176 eV.87

Warning: Comparing calculated and measured binding energies is an important but not definitive way to
assess a model-based wave function.
The energy (4.D.7a) is an integral over all space; its integrand emphasizes certain regions of space more than
others. So agreement of calculated and measured energies does not mean that the model-based wave function is
correct. For this reason it is important to examine other physically meaningful properties one can calculate from
wave functions. If such quantities cannot easily be measured, we can compare to values from more accurate
(and difficult) theoretical calculations. To illustrate, Ill compare a couple of radial expectation values
calculated using my model-based radial function against values calculated using highly accurate Hartree-Fock
radial functions. The latter functions are also based on the pure-Coulomb Hamiltonian [Eq. (4.D.1d), p. 338]
but take into account the electron-electron interactions and electron spin (see Complement 11.B of Chap. 11).88
Table 4.D.1 compares the expectation value of r and of r2 for a 2s electron using equations in 4.3.5. The
values from our model potential energy agree quite well with the Hartree-Fock results and are close to the
experimentally determined atomic size 3.15 a0 = 1.67
A.
Such excellent agreement, along with the agreement of the bound-state energy E2s with experiment, strongly
suggests but does not prove that my model-based 2s radial function is accurate. The three numbers E2s , hri2s ,
and hr2 i2s , depend on my function in different regions of r. To clarify this point, Fig. 4.D.5, p. 344 compares

87 Jargon: Experimentalists refer to the binding energy of the valence electron in an atom as its ionization potential (or
ionization energy ). Sometimes the term binding energy is used to refer only to so-called inner-shell electrons, those electrons
localized nearest the nucleus. Ill follow the most common usage: binding energy refers to any electron, and ionization
potential refers to the amount of energy required to remove the valence electron to infinity.
88 Read on: More precisely, these calculations use the Hartree-Fock self-consistent field method. This approach is based on
the independent particle model (Chap. 11). For a preview, see Chap. 10 of Quantum States of Atoms, Molecules, and Solids
by myself, Thomas L. Estle, and Neal. F. Lane (Englewood Cliffs, New Jersey: Prentice-Hall, 1977) For details, see the original
reference: E. Clementi, Atomic Data Nuclear Data Tables 14, 177, (1974). I use a Hartree-Fock program for atoms written
by R. D. Cowan. You can access an on-line applet based on this program at http://plasma-gate.weizmann.ac.il/~fnralch/
rcn.html. This site also links to one where you can download Cowans program, which is in FORTRAN.

JQPMaster

Version: 8.35

Printed: August 17, 2010

344

4.D.4 A radial function for the valence electron in Li


Table 4.D.1.
Radial mean values for the 2s state of atomic
lithium. The last column shows results of a Hartree-Fock calculation.
For our present purposes, these results
constitute the answer in the back of
the book against which results from
our model should be compared.

quantity

screened Coulomb model

hri2s (a0 )

Hartree-Fock value

3.943

hr2 i2s (a20 )

3.874

17.79

17.74

the integrands in the mean values of r and r2 : the expectation value hr2 i2s gives more weight to values of r > 1
than does hri2s .89
J

Figure 4.D.5. Integrands of the radial integrals for expectation values.


Shown are integrands for r (solid curve)
and r2 (dash curve) for the 2s electron
in Li, as calculated from the model radial
function Eq. (4.D.8).

Try This!

integrands in expectation values

in Xr 2 \ 2 s
2

in Xr\ 2 s

10

r Ha0 L

4.11. How certain can we be about the position of the 2s electron?


Evaluate (r)2s from the data in Tbl. 4.D.1. How accurately does our simple model of Li predict this
important quantity? What does the value of this uncertainty tell you about the 2s electron in Li? Would
you expect (r)1s to be less than, equal to, or greater than (r)2s ? Why?

Try This!

4.12. Always more details. . .


Explain how to get from Eq. (4.D.7a) to Eq. (4.D.7b).

Aside. How big is a lithium atom?


Physicists and chemists often refer to the shell radius of an atom. By this they usually mean the value of r
rad
rad
at which Pn,`
(r) attains its maximum value. For the 2s electron in Li, P2s
(r) peaks at r = 3.0 a0 . But the
shell radius is not always the most meaningful measure of the size of an atom. Figure 4.D.3 shows that the
2s electron is more likely to be found in a shell at 3.0 a0 than anywhere else, but this number alone doesnt
adequately convey the electrons position properties. Arguably, the mean value hri2s = 3.943 a0 , is a more
sensible estimate of the size of this state.

Aside. The radial function for the model potential energy of Eq. (4.D.4a).
The approximate asymptotic solution of the radial equation for the model potential energy Vval of Eq. (4.D.4a)
for a neutral atom, in atomic units (Appendix F) is90

1 1 r
`(` + 1)
1
1

e
Rn,` (r) = A r
1 2
1 +
(for r 1/),
(4.D.8)
2 r
2r

89 Commentary: This example illustrates the difficulty in assessing a wave function by calculating quantities that require
integration over space: the integrand may emphasize the wave function in one region of space over another. The key point
is that, just because your model can produce one property accurately, doesnt mean that it can predict other quantities with
comparable accuracy. This issue becomes critical in applications of the variational method (Chap. 15).
90 Read on: Note the qualifier: this approximate function is valid only at large r, and it neglects terms of order r 2 and
higher. Although it doesnt blow up at the origin, it may predict the wrong number of nodes if used at smaller r. For more
information (and tables of parameters for lots of atoms), see Table 4.2 in Radzig and Smirnov (1985).

JQPMaster

Version: 8.35

Printed: August 17, 2010

345
where, for the 2s electron (` = 0 ) in Li, the parameters are A = 0.82 and = 0.630.91
Key Points
When confronted with a complicated, unfamiliar system, use The Principle of Modeling to approximate the
system by models that belong to a class of problems you know how to solve.
The (spherically symmetric) screened Coulomb model potential energy, together with the assumption
of infinite nuclear mass, can reduce the complicated non-central problem of the valence electron in an alkali
atom to a central problem of a single electron.
rad
The shell probability provides a direct physical picture of Pn,`
(r).

Complement 4.E. How to cope with an arbitrary wave function


He had not liked what they said of matter. He had not liked what they said of the world.
They had taken the difficulties in the fundamental theory of matter as a license to distort
the nature of the real, so that Justin could not help but perceive their views as acts of
ontological sabotage: maligning the objectivity of matter and unraveling the rationality of
the world.
Properties of Light, by Rebecca Goldstein

Throughout most of this book well focus on stationary angular momentum eigenstates (SAMEs) of rotationally
invariant systems. But the ideas and tools we are developing can be used to study more general states. The technique
that makes this possible is one of the most widely used in all of quantum mechanics: the method of eigenfunction
expansion (Appendix A). In this Complement, Ill show you how to apply this method to 3D SAMEs.
4.E.1 The method of eigenfunction expansion
In a nutshell, this method tells us to expand the function we dont know in a complete set of functions we do know. For a
rotationally invariant 3D system, we expand (r, t) in the complete set of SAMEs, { E,`,m` (r) }. This set is complete
b b
because it consists of mathematically distinct eigenfunctions of a complete set of commuting operators, { H,
L 2, b
Lz }.
Because these eigenfunctions depend on (r, , ), we can use { E,`,m` (r) } as the basis for expansion of any function
of r that obeys the same boundary conditions as E,`,m` (r). If the function depends on additional independent
variables, as (r, t) depends on t, then that dependence must appear in the expansion coefficients.
For convenience, Ill discuss a system that supports only bound states (for example, a 3D simple harmonic
oscillator). We expand an arbitrary wave function of such a system as
(r, t) =

cn,`,m` (0) n,`,m` (r) ei En,` t/~ .

(4.E.1a)

n,`,m`

The argument (0) on the coefficients cn,`,m` (0) reminds us that these coefficients are determined by the initial wave
function, (r, 0). In Dirac and conventional notation, the expansion coefficients are
Z
cn,`,m` (0) = hn,`,m` | (0)ir =

2
0

n,`,m
(r) (r, 0) r2 dr sin d d
`

(4.E.1b)

91 Commentary: The quantity is the decay constant of the 2s electron [Eq. (5.4.5b), p. 367]. [This quantity is determined
from the ionization potential of the electron, the (positive) amount of energy required to ionize the the electron to an infinite
distance from the nucleus. For Li, the experimental ionization potential is 5.39176 eV.] The appearance of er in Eq. (4.D.8)
and its dominance as r is a major plus for this model. The function Rn,` (r) of (4.D.8) is strictly valid only for r 1,
well outside the core where the two 1s electrons hang out. But, as discussed in 4.7, the core of a lithium atom is quite
small, so Eq. (4.D.8) is an excellent approximation for all r, and we can use it to calculate properties like hri2s and (r)2s .
The normalization integral for this function, calculated via numerical integration, turns out to be 1.03544. The slight deviation
from 1 is due to deficiencies in this function at very small values of r.

JQPMaster

Version: 8.35

Printed: August 17, 2010

346

4.E.2 Calculation of expectation values and related quantities

Given (r, 0), we can calculate the coefficients from Eq. (4.E.1b), then plug them into Eq. (4.E.1a) to get the wave
function for any t 0. But we must ensure that (r, t) is normalized by imposing the condition
h(0) | (0)i = 1 =

|cn,`,m` (0)|2 = 1

normalization condition (3D)

(4.E.1c)

n,`,m`

Tips for application of the method of eigenfunction expansion.


(1) The sum over the principal quantum number n must include all stationary-state energies.92 If the system
supports an infinite number of bound states, then nmax = , and the number of terms required by Eq. (4.E.1a),
in principle, is infinite.93 In practice, we often truncate the sum over n at just a few termsthat is, we impose
an upper limit nmax on this sum. How large nmax must be depends on how accurate our results must be.
(2) The sum over the orbital quantum number ` must include all allowed values of `. For some systems ` can be
any non-negative integer, ` = 0, 1, . . . , in which case this sum runs from ` = 0 to `max = . For others, `
has a maximum finite value determined by n. For example, in the pure-Coulomb model of atomic hydrogen
(Chap. 5), `max = n 1. In any case, we can usually truncate the sum over ` with little loss of accuracy.
(3) The sum over the projection quantum number m` must run over the (2` + 1) allowed values m` = `, . . . , +`.

Interpreting the expansion coefficients. The coefficients { cn,`,m` (0) } derive physical significance from the
generalized Born interpretation . I think about the eigenfunction expansion (4.E.1a) as a device that translates
information about the state from the language of position, as expressed by (r, 0), to the language of the observables
whose operators appear in the CSCO. In that language, the information is contained in the expansion coefficients
{ cn,`,m` (0) }:
Rule: The coefficient cn,`,m` (0) is a probability amplitude for simultaneous measurement (at t = 0) of all the observables
b b
in { H,
L 2, b
Lz }the energy, the magnitude of the orbital angular momentum, and the projection of L on the axis of spatial
quantization:94

the probability of obtaining the values E = En,` and

2
L = ~2 `(` + 1) and Lz = m` ~ in a simultaneous
2
|cn,`,m` (0)| =
(4.E.2)

measurement of energy and the magnitude of L and


Lz at t = 0.

4.E.2 Calculation of expectation values and related quantities

Having obtained the coefficients { cn,`,m` (0) }, we can easily calculate expectation values and uncertainties of various
observables. Of course, in many experiments we dont measure all the observables represented by the CSCO, so we
have to keep in mind that
Rule: When calculating expectation values, sum probabilities (not probability amplitudes) over all unmeasured observables.

Example 4.19 (The mean value and uncertainty of Lz in an arbitrary state.)


To evaluate the average value of Lz at t = 0 from the probability amplitudes cn,`,m` (0), I must multiply each
possible measurement outcome, m` ~, by a weighting factor equal to the probability of obtaining that outcome.
If only Lz is measured, the unmeasured observables that define the SAMEs are the energy and the magnitude
of the L. So my weighting factor must include sums over the corresponding quantum numbers:

92 Memory Jog: For systems that support both bound and continuum states. this sum consists of two terms. The first
is a sum over the bound-state energies; the second is an integral over the continuum energies. If the contributions of the
continuum-energy terms are negligible, then the analysis and interpretation reduces to that of this section. Be careful, though.
In Eliminating the continuum states can lead to serious errors. For example, in the collision of electrons with hydrogen atoms,
the continuum makes a major contribution to the cross sections.
93 Commentary: For example, a 3D harmonic oscillator potential is infinite and therefore supports an infinite number of bound
states. But a potential need not be infinite to have this property: its also true of the pure Coulomb potential (Chap. 5).
94 Commentary: The coefficients c
n,`,m` (0) may provide information about other observables, such as the parity. But they
dont tell us anything directly about position or momentum; for that, we should turn to (r, 0) or to its momentum-space
counterpart (the Fourier transform).
JQPMaster

Version: 8.35

Printed: August 17, 2010

4.E.2 Calculation of expectation values and related quantities

hLz i(0) =

"n
max `X
max
X

`
X
m` =`

n=0

`=0

347
#
2

|cn,`,m` (0)|
{z

m` ~.

(4.E.3a)

weighting factor

To find the corresponding uncertainty in Lz for this state,


p
Lz (0) = hL2z i(0) [hLz i(0)]2 ,

(4.E.3b)

I use (4.E.3a) for the second number under the square root and, for the first number, the expectation value
of L2z ,
"n
#
`
max `X
max
X
X
2
2
|cn,`,m` (0)|
(m` ~)2 .
hLz i(0) =
(4.E.3c)
m` =`

n=0 `=0

Given an initial wave function (r, 0) and V (r), I can use Eqs. (4.E.3) to evaluate statistical quantities at
t = 0 (or, for that matter, for any t 0).
J

b 2 but not of L
bz .)
Example 4.20 (An eigenstate of L
Suppose, instead of a completely general initial state, I had chosen (r, 0) in Example 4.19 to be an eigenfunction of b
L 2 with eigenvalue ~2 `0 (`0 + 1). How would the results change? The functions in { n,`,m` (r) }
are linearly independent, so if (r, 0) is an eigenfunction of b
L 2 with this eigenvalue, then each term on the
right-hand side of Eq. (4.E.1a) must be an eigenfunction with that eigenvalue:95
#
"n
`0
max
X
X
i En,`0 t/~
cn,`0 ,m` n,`0 ,m` (r) e
.
`0 (r, t) =
(4.E.4)
m` =`0

n=0

Rule: If the function being expanded represents a state in which an observable in the CSCO is sharp, then dont sum
over the corresponding quantum number.
In this example, I didnt sum over `, because our state prescribes a value of this observable.

95 Details: I could have derived this result from Eq. (4.E.1b). Since only eigenfunctions of b
L 2 with orbital quantum number `0
appear in the sum over `, the expansion coefficient cn,`,m` (0) in Eq. (4.E.1a) collapse to cn,`,m` (0) = cn,`0 ,m` `,`0 , where the
Kronecker delta `,`0 ensures that no unwanted spatial functions creep into the sum.

JQPMaster

Version: 8.35

Printed: August 17, 2010

Anda mungkin juga menyukai