Numerical Methods in Quantum Mechanics by Paolo Giannozzi - HTML preview

PLEASE NOTE: This is an HTML preview only and some elements such as links or page numbers may be incorrect.
Download the book in PDF, ePub, Kindle for a complete version.

+

r2 (E − V (r)) −

+

y(x) = 0

(2.35)

dx2

¯

h2

2

where V (r) = −Zq2e/r for the Coulomb potential. This equation no longer

presents any singularity for r = 0, is in the form of Eq.(1.21), with 2m

2

e

1

g(x) =

r(x)2 (E − V (r(x))) −

+

(2.36)

¯

h2

2

and can be directly solved using the numerical integration formulae Eqs.(1.31)

and Eq.(1.32) and an algorithm very similar to the one of Sect.1.3.2.

Subroutine do mesh defines at the beginning and once for all the values of

r,

r, r2 for each grid point. The potential is also calculated once and for all

in init pot. The grid is calculated starting from a minimum value x = −8,

corresponding to Zrmin

3.4 × 10−3 Bohr radii. Note that the grid in r does

not include r = 0: this would correspond to x = −∞. The known analytical

behavior for r → 0 and r → ∞ are used to start the outward and inward

recurrences, respectively.

2.3.2

Improving convergence with perturbation theory

A few words are in order to explain this section of the code:

i = icl

ycusp = (y(i-1)*f(i-1)+f(i+1)*y(i+1)+10.d0*f(i)*y(i)) / 12.d0

dfcusp = f(i)*(y(i)/ycusp - 1.d0)

! eigenvalue update using perturbation theory

de = dfcusp/ddx12 * ycusp*ycusp * dx

22

whose goal is to give an estimate, to first order in perturbation theory, of the

difference δe between the current estimate of the eigenvalue and its final value.

Reminder: icl is the index corresponding to the classical inversion point.

Integration is made with forward recursion up to this index, with backward

recursion down to this index. icl is thus the index of the matching point

between the two functions. The function at the right is rescaled so that the

total function is continuous, but the first derivative dy/dx will be in general

discontinuous, unless we have reached a good eigenvalue.

In the section of the code shown above, y(icl) is the value given by Nu-

merov’s method using either icl-1 or icl+1 as central point; ycusp is the value

predicted by the Numerov’s method using icl as central point. The problem

is that ycusp=y(icl).

What about if our function is the exact solution, but for a different problem?

It is easy to find what the different problem could be: one in which a delta func-

tion, v0δ(x−xc), is superimposed at xc ≡x(icl) to the potential. The presence

of a delta function causes a discontinuity (a ”cusp”) in the first derivative, as

can be demonstrated by a limit procedure, and the size of the discontinuity is

related to the coefficient of the delta function. Once the latter is known, we

can give an estimate, based on perturbation theory, of the difference between

the current eigenvalue (for the ”different” potential) and the eigenvalue for the

”true” potential.

One may wonder how to deal with a delta function in numerical integration.

In practice, we assume the delta to have a value only in the interval ∆x centered

on y(icl). The algorithm used to estimate its value is quite sophisticated. Let

us look again at Numerov’s formula (1.32): note that the formula actually provides only the product y(icl)f(icl). From this we usually extract y(icl)

since f(icl) is assumed to be known. Now we suppose that f(icl) has a

different and unknown value fcusp, such that our function satisfies Numerov’s

formula also in point icl. The following must hold:

fcusp*ycusp = f(icl)*y(icl)

since this product is provided by Numerov’s method (by integrating from icl-1

to icl+1), and ycusp is that value that the function y must have in order to

satisfy Numerov’s formula also in icl. As a consequence, the value of dfcusp

calculated by the program is just fcusp-f(icl), or δf .

The next step is to calculate the variation δV of the potential V (r) appearing

in Eq.(2.35) corresponding to δf . By differentiating Eq.(2.36) one finds δg(x) =

−(2me/¯h2)r2δV . Since f (x) = 1 + g(x)(∆x)2/12, we have δg = 12/(∆x)2δf ,

and thus

¯

h2 1

12

δV = −

δf.

(2.37)

2me r2 (∆x)2

First-order perturbation theory gives then the corresponding variation of the

eigenvalue:

δe = ψ|δV |ψ =

|y(x)|2r(x)2δV dx.

(2.38)

23

Note that the change of integration variable from dr to dx adds a factor r to

the integral:

dr(x)

f (r)dr =

f (r(x))

dx =

f (r(x))r(x)dx.

(2.39)

0

−∞

dx

−∞

We write the above integral as a finite sum over grid points, with a single

non-zero contribution coming from the region of width ∆x centered at point

xc =x(icl). Finally:

¯

h2

12

δe = |y(xc)|2r(xc)2δV ∆x = −

|y(xc)|2∆xδf

(2.40)

2me (∆x)2

is the difference between the eigenvalue of the current potential (i.e. with a

superimposed Delta function) and that of the true potential. This expression

is used by the code to calculate the correction de to the eigenvalue. Since in

the first step this estimate may have large errors, the line

e = max(min(e+de,eup),elw)

prevents the usage of a new energy estimate outside the bounds [elw,eip]. As

the code proceeds towards convergence, the estimate becomes better and better

and convergence is very fast in the final steps.

2.3.3

Laboratory

• Examine solutions as a function of n and ; verify the presence of acci-

dental degeneracy.

• Examine solutions as a function of the nuclear charge Z.

• Compare the numerical solution with the exact solution, Eq.(2.29), for the 1s case (or other cases if you know the analytic solution).

• Slightly modify the potential as defined in subroutine init pot, verify

that the accidental degeneracy disappears. Some suggestions: V (r) =

−Zq2e/r1+δ where δ is a small, positive or negative, number; or add an

exponential damping (Yukawa) V (r) = −Zq2e exp(−Qr)/r where Q is a

number of the order of 0.05 a.u..

• Calculate the expectation values of r and of 1/r, compare them with the

known analytical results.

Possible code modifications and extensions:

• Consider a different mapping: r(x) = r0(exp(x) − 1), that unlike the one

we have considered, includes r = 0. Which changes must be done in order

to adapt the code to this mapping?

24

Chapter 3

Scattering from a potential

Until now we have considered the discrete energy levels of simple, one-electron

Hamiltonians, corresponding to bound, localized states. Unbound, delocalized

states exist as well for any physical potential (with the exception of idealized

models like the harmonic potential) at sufficiently high energies. These states

are relevant in the description of elastic scattering from a potential, i.e. pro-

cesses of diffusion of an incoming particle. Scattering is a really important

subject in physics, since what many experiments measure is how a particle is

deflected by another. The comparison of measurements with calculated results

makes it possible to understand the form of the interaction potential between

the particles. In the following a very short reminder of scattering theory is

provided; then an application to a real problem (scattering of H atoms by rare

gas atoms) is presented. This chapter is inspired to Ch.2 (pp.14-29) of the book

of Thijssen.

3.1

Short reminder of the theory of scattering

The elastic scattering of a particle by another is first mapped onto the equivalent

problem of elastic scattering from a fixed center, using the same coordinate

change as described in the Appendix. In the typical geometry, a free particle,

described as a plane wave with wave vector along the z axis, is incident on the

center and is scattered as a spherical wave at large values of r (distance from the

center). A typical measured quantity is the differential cross section dσ(Ω)/dΩ,

i.e. the probability that in the unit of time a particle crosses the surface in the

surface element dS = r2dΩ (where Ω is the solid angle, dΩ = sin θdθdφ, where

θ is the polar angle and φ the azimuthal angle). Another useful quantity is the

total cross section σtot = (dσ(Ω)/dΩ)dΩ. For a central potential, the system

is symmetric around the z axis and thus the differential cross section does not

depend upon φ. The cross section depends upon the energy of the incident

particle.

Let us consider a solution having the form:

f (θ)

ψ(r) = eik·r +

eikr

(3.1)

r

25

with k = (0, 0, k), parallel to the z axis. This solution represents an incident

plane wave plus a diffused spherical wave. It can be shown that the cross section

is related to the scattering amplitude f (θ):

dσ(Ω) dΩ = |f(θ)|2dΩ = |f(θ)|2 sin θdθdφ.

(3.2)

dΩ

Let us look for solutions of the form (3.1). The wave function is in general given by the following expression:

l

χl(r)

ψ(r) =

Alm

Ylm(θ, φ),

(3.3)

r

l=0 m=−l

which in our case, given the symmetry of the problem, can be simplified as

χl(r)

ψ(r) =

Al

Pl(cos θ).

(3.4)

r

l=0

The functions χl(r) are solutions for (positive) energy E = ¯

h2k2/2m with an-

gular momentum l of the radial Schrödinger equation:

¯

h2 d2χl(r)

¯

h2l(l + 1)

+ E − V (r) −

χl(r) = 0.

(3.5)

2m

dr2

2mr2

It can be shown that the asymptotic behavior at large r of χl(r) is

χl(r)

kr [jl(kr) cos δl − nl(kr) sin δl]

(3.6)

where the jl and nl functions are the well-known spherical Bessel functions,

respectively regular and divergent at the origin. These are the solutions of the

radial equation for Rl(r) = χl(r)/r in absence of the potential. The quantities

δl are known as phase shifts. We then look for coefficients of Eq.(3.4) that yield the desired asymptotic behavior (3.1). It can be demonstrated that the cross section can be written in terms of the phase shifts. In particular, the differential

cross section can be written as

2

1

=

(2l + 1)eiδl sin δlPl(cos θ) ,

(3.7)

dΩ

k2 l=0

while the total cross section is given by

4π ∞

2mE

σtot =

(2l + 1) sin2 δl

k =

.

(3.8)

k2

¯

h2

l=0

Note that the phase shifts depend upon the energy. The phase shifts thus con-

tain all the information needed to know the scattering properties of a potential.

26

index-31_1.jpg

3.2

Scattering of H atoms from rare gases

The total cross section σtot(E) for the scattering

of H atoms by rare gas atoms was measured by

Toennies et al., J. Chem. Phys. 71, 614 (1979).

At the right, the cross section for the H-Kr sys-

tem as a function of the energy of the center of

mass. One can notice “spikes” in the cross sec-

tion, known as “resonances”. One can connect

the different resonances to specific values of the

angular momentum l.

The H-Kr interaction potential can be modelled quite accurately as a Lennard-

Jones (LJ) potential:

σ 12

σ 6

V (r) =

− 2

(3.9)

r

r

where

= 5.9meV, σ = 3.57˚

A. The LJ potential is much used in molecular and

solid-state physics to model interatomic interaction forces. The attractive r−6

term describes weak van der Waals (or “dispersive”, in chemical parlance) forces

due to (dipole-induced dipole) interactions. The repulsive r−12 term models the

repulsion between closed shells. While usually dominated by direct electrostatic

interactions, the ubiquitous van der Waals forces are the dominant term for the

interactions between closed-shell atoms and molecules. These play an important

role in molecular crystals and in macro-molecules. The LJ potential is the first

realistic interatomic potential for which a molecular-dynamics simulation was

performed (Rahman, 1965, liquid Ar).

It is straightforward to find that the LJ potential as written in (3.9) has a minimum Vmin = − for r = σ, is zero for r = σ/21/6 = 0.89σ and becomes

strongly positive (i.e. repulsive) at small r.

3.2.1

Derivation of Van der Waals interaction

The Van der Waals attractive interaction can be described in semiclassical terms

as a dipole-induced dipole interaction, where the dipole is produced by a charge

fluctuation. A more quantitative and satisfying description requires a quantum-

mechanical approach. Let us consider the simplest case: two nuclei, located

in RA and RB, and two electrons described by coordinates r1 and r2. The

Hamiltonian for the system can be written as

¯

h2

q2

¯

h2

q2

H

=

∇2 −

e

∇2 −

e

(3.10)

2m

1

|r

2

1 − RA|

2m

|r2 − RB|

q2

q2

q2

q2

e

e

+

e

+

e

,

|r1 − RB|

|r2 − RA|

|r1 − r2|

|RA − RB|

where ∇i indicates derivation with respect to variable ri, i = 1, 2. Even this

“simple” hamiltonian is a really complex object, whose general solution will be

the subject of several chapters of these notes. We shall however concentrate on

27

the limit case of two Hydrogen atoms at a large distance R, with R = RA −RB.

Let us introduce the variables x1 = r1−RA, x2 = r2−RB. In terms of these new

variables, we have H = HA + HB + ∆H(R), where HA (HB) is the Hamiltonian

for a Hydrogen atom located in RA (RB), and ∆H has the role of “perturbing

potential”:

q2

q2

q2

q2

∆H = −

e

e

+

e

+ e .

(3.11)

|x1 + R|

|x2 − R|

|x1 − x2 + R|

R

Let us expand the perturbation in powers of 1/R. The following expansion,

valid for R → ∞, is useful:

1

1

R · x

3(R · x)2 − x2R2

+

.

(3.12)

|x + R|

R

R3

R5

Using such expansion, it can be shown that the lowest nonzero term is

2q2

(R · x1)(R · x2)

∆H

e

x1 · x2 − 3

.

(3.13)

R3

R2

The problem can now be solved using perturbation theory. The unperturbed

ground-state wavefunction can be written as a product of 1s states centered

around each nucleus: Φ0(x1, x2) = ψ1s(x1)ψ1s(x2) (it should actually be anti-

symmetrized but in the limit of large separation it makes no difference.). It is

straightforward to show that the first-order correction, ∆(1)E = Φ0|∆H|Φ0 ,

to the energy, vanishes because the ground state is even with respect to both x1

and x2. The first nonzero contribution to the energy comes thus from second-

order perturbation theory:

| Φi|∆H|Φ0 |2

∆(2)E = −

(3.14)

E

i>0

i − E0

where Φi are excited states and Ei the corresponding energies for the unper-

turbed system. Since ∆H ∝ R−3, it follows that ∆(2)E = −C6/R6, the well-

known behavior of the Van der Waals interaction.1 The value of the so-called C6 coefficient can be shown, by inspection of Eq.(3.14), to be related to the product of the polarizabilities of the two atoms.

3.3

Code: crossection

Code crossection.f902, or crossection.c3, calculates the total cross section σtot(E) and its decomposition into contributions for the various values of the

angular momentum for a scattering problem as the one described before.

The code is composed of different parts. It is always a good habit to verify

separately the correct behavior of each piece before assembling them into the

final code (this is how professional software is tested, by the way). The various

parts are:

1Note however that at very large distances a correct electrodynamical treatment yields a

different behavior: ∆E ∝ −R−7.

2http://www.fisica.uniud.it/%7Egiannozz/Corsi/MQ/Software/F90/crossection.f90

3http://www.fisica.uniud.it/%7Egiannozz/Corsi/MQ/Software/C/crossection.c

28

1. Solution of the radial Schrödinger equation, Eq.(3.5), with the Lennard-Jones potential (3.9) for scattering states (i.e. with positive energy). One can simply use Numerov’s method with a uniform grid and outwards

integration only (there is no danger of numerical instabilities, since the

solution is oscillating). One has however to be careful and to avoid the

divergence (as ∼ r−12) for r → 0. A simple way to avoid running into

trouble is to start the integration not from rmin = 0 but from a small but

nonzero value. A suitable value might be rmin ∼ 0.5σ, where the wave

function is very small but not too close to zero. The first two points can be

calculated in the following way, by assuming the asymptotic (vanishing)

form for r → 0:

2m σ12

2m σ12

χ (r)

χ(r)

=⇒

χ(r)

exp

r−5

(3.15)

¯

h2 r12

25¯

h2

(note that this procedure is not very accurate because it introduces and

error of lower order, i.e. worse, than that of the Numerov’s algorithm.

In fact by assuming such form for the first two steps of the recursion,

we use a solution that is neither analytically exact, nor consistent with

Numerov’s algorithm).

The choice of the units in this code is (once again) different from that of

the previous codes. It is convenient to choose units in which ¯

h2/2m is a

number of the order of 1. Two possible choices are meV/˚

A2, or meV/σ2.

This code uses the former choice. Note that m here is not the electron

mass! it is the reduced mass of the H-Kr system. As a first approximation,

m is here the mass of H.

2. Calculation of the phase shifts δl(E). Phase shifts can be calculated by

comparing the calculated wave functions with the asymptotic solution at

two different values r1 and r2, both larger than the distance rmax beyond

which the potential can be considered to be negligible. Let us write

χl(r1) = Akr1 [jl(kr1) cos δl − nl(kr1) sin δl]

(3.16)

χl(r2) = Akr2 [jl(kr2) cos δl − nl(kr2) sin δl] ,

(3.17)

from which, by dividing the two relations, we obtain an auxiliary quantity

K

r2χl(r1)

jl(kr1) − nl(kr1) tan δl

K ≡

=

(3.18)

r1χl(r2)

jl(kr2) − nl(kr2) tan δl

from which we can deduce the phase shift:

Kjl(kr2) − jl(kr1)

tan δl =

.

(3.19)

Knl(kr2) − nl(kr1)

The choice of r1 and r2 is not very critical. A good choice for r1 is about

at rmax. Since the LJ potential decays quite quickly, a good choice is

r1 = rmax = 5σ. For r2, the choice is r2 = r1 + λ/2, where λ = 1/k is the

wave length of the scattered wave function.

29

3. Calculation of the spherical Bessel functions jl and nl. The analytical

forms of these functions are known, but they become quickly unwieldy for

high l. One can use recurrence relation. In the code the following simple

recurrence is used:

2l + 1

zl+1(x) =

zl(x) − zl−1(x),

z = j, n

(3.20)

x

with the initial conditions

cos x

sin x

sin x

cos x

j−1(x) =

,

j0(x) =

;

n−1(x) =

,

n0(x) = −

.

x

x

x

x

(3.21)

Note that this recurrence is unstable for large values of l: in addition to

the good solution, it has a ”bad” divergent solution. This is not a serious

problem in practice: the above recurrence should be sufficiently stable up

to at least l = 20 or so, but you may want to verify this.

4. Sum the phase shifts as in Eq.(3.8) to obtain the total cross section and a graph of it as a function of the incident energy. The relevant range of

energies is of the order of a few meV, something like 0.1 ≤ E ≤ 3 ÷ 4

meV. If you go too close to E = 0, you will run into numerical difficulties

(the wave length of the wave function diverges). The angular momentum

ranges from 0 to a value of lmax to be determined.

The code writes a file containing the total cross section and each angular mo-

mentum contribution as a function of the energy (beware: up to lmax = 13; for

larger l lines will be wrapped).

3.3.1

Laboratory

• Verify the effect of all the parameters of the calculation: grid step for in-

tegration, rmin, r1 = rmax, r2, lmax. In particular the latter is important:

you should start to find good results when lmax ≥ 6.

• Print, for some selecvted values of the energy, the wave function and its

asymptotic behavior. You should verify that they match beyond rmax.