Anda di halaman 1dari 24

Physics 125c

Course Notes
Approximate Methods
Solutions to Problems
040415 F. Porter

1 Exercises
1. Prove the theorem quoted in section ??:
Theorem: If we have a normalized function |ψi such that
E0 ≤ hψ|H|ψi ≤ E1 , (1)
then
hHψ|Hψi − hψ|H|ψi2
E0 ≥ hψ|H|ψi − . (2)
E1 − hψ|H|ψi
Solution: The theorem is equivalent to the statement
hψ|H 2 ψi − hψ|H|ψi2 ≥ (hψ|H|ψi − E0 )(E1 − hψ|H|ψi). (3)
Notice that if we add a constant A to H, obtaining H ′ = H + A (hence
also En → En′ = En + A), both sides of this inequality are unaltered.
The left hand side is a measure of the width of the energy distribution
and is not altered by shifting the energy scale. Likewise, the right hand
side only depends on energy differences. Thus, as long as the spectrum
of H is bounded below, the problem is equivalent to a problem where
the spectrum is non-negative. In particular, we may simplify by taking
E0 = 0.
Hence, consider:
hψ|H 2ψi − hψ|H|ψi2 − hψ|H|ψi(E1 − hψ|H|ψi)
= hψ|H 2ψi − E1 hψ|H|ψi (4)
|cn |2 En2 − E1 |cn |2 En
X X
= (5)
n=0 n=0
2
X
= |cn | En (En − E1 ) (6)
n=0
|cn |2 En (En − E1 ),
X
= since E0 = 0, (7)
n=1
≥ 0, since each term in the sum is non-negative.

1
2. Let us pursue our variational approach to the estimation of ground state
energy levels of atoms to the “general” case. We consider an atom with
nuclear charge Z, and N electrons. The Hamiltonian of interest is:
H(Z, N) = Hkin − ZVc + Ve (8)
N
X p2n
Hkin = , (9)
n=1 2m
where (10)
N
X 1
Vc = α (11)
n=1 |xn |
X 1
Ve = α (12)
N ≥j>k≥1|xk − xj |
m = electron mass (13)
α = fine structure constant. (14)
Denote the ground state energy of H(Z, N) by −B(Z, N), with B(Z, 0) =
0.

(a) Generalize the variational calculation we performed for the ground


state of helium to the general Hamiltonian H(Z, N). Thus, se-
lect your “trial function” to be a product of N identical “hydro-
gen atom ground state” functions. Determine the resulting lower
bound B̂(Z, N) on B(Z, N) (i.e., an upper limit on the ground
state energies).
Solution: Let Ze be the fixed nuclear charge in the Hamiltonian,
and let z be the effective Z variational parameter. The trial wave
function we are told to use is thus:
s
N
Y z 3 − az (rn ) 1
ψZN (x1 , . . . , xN ) = e 0 , a0 = . (15)
n=1 πa30 mα
The expectation value of the total kinetic energy for this trial
function is:
p2 1
Hkin = Nhψ| 1 |ψi = Nz 2 mα2 . (16)
2m 2
The expectation value of the potential energy of the electrons in
the nuclear electric field is:
−ZVc = −NZzmα2 . (17)

2
The expectation value of the potential energy of the electrons in
the fields of the other electrons is:
N(N − 1) 1 5
Ve = z mα2 . (18)
2 24
Putting these terms together, we have
1 5
 
hH(Z, N)iz = mα2 Nz z − 2Z + (N − 1) (19)
2 8
We minimize with respect to z:
d 2 5 5
 
0= z − 2Zz + (N − 1)z = 2z − 2Z + (N − 1). (20)
dz 8 8
Thus, the minimum occurs at
5
z=Z− (N − 1) (21)
16
The variational bound on the (negative of the) ground state ener-
gies is then:
2
1 5
 
2
B̂(Z, N) = −hH(Z, N)imin = mα N Z − (N − 1) . (22)
2 16
(b) Make a simple table comparing your variational bounds with the
observed ground state energies for lithium, beryllium, and nitro-
gen. Note that a simple web search for “ionization potentials” will
get you a multitude of tables of observed values, or you can look
at a reference such as the CRC Press’s Handbook of Chemistry
and Physics. The table entries are typically of the form:
B(Z, N) − B(Z, N − 1).

Solution: A Google search on “ionization potentials” results in


many suitable hits, including:
http://www.chemistrycoach.com/ionization potentials f.htm
The ionization potentials for lithium, beryllium, and nitrogen are
reproduced in Table 2b.

The predicted bounds on the energies, according to Eqn. 22, are


compared with the observed values in Table 2b.

3
Ionization Potentials for Lithium, Beryllium, and Nitrogen
(from http://www.chemistrycoach.com/ionization potentials f.htm)
Lithium Beryllium Nitrogen
1st I.P. 5.4 9.3 14.5
2nd I.P. 75.6 18.2 29.6
3rd I.P. 122 154 47.5
4th I.P. 218 77.5
5th I.P. 97.9
6th I.P. 552
7th I.P. 667

Comparison of Variational Prediction with Measured Energies


Lithium Beryllium Nitrogen
N Pred. Meas. Pred. Meas. Pred. Meas.
1 122.5 122 217.7 218 666.7 667
2 196.5 198 370.0 372 1217.0 1219
3 230.2 203 464.9 390 1658.8 1317
4 510.4 400 2000.2 1394
5 2249.2 1442
6 2413.6 1472
7 2501.5 1486

4
(c) Do your results make sense? If not, can you figure out what is
wrong, and whether the calculation we did for He is to be trusted?

Solution: Note that for N = 1, the calculation is “exact”, up to the


approximations made and effects neglected. The differences between
predicted and observed values for N = 1 are an indication of the un-
certainty due to the neglect in these matters, and possibly experimental
uncertainties.
We see that for N > 2, the computed bounds are always violated by the
data. The trial wave function for N > 2 is not properly antisymmetric
under interchange of the electrons, hence is not in the Hilbert space.
The calculation for He is still all right, since the electron has a spin
degree of freedom – A symmetric spatial wave function for two electrons
is permitted, as the spin wave function can be antisymmetric.
The computation for N = 2 should be all right, and we see that the
bound is always on the side it is supposed to be. Indeed, we do rather
well, always getting within a percent of the actual energy.
3. We consider the quantum mechanics of a particle in the earth’s gravi-
tational field:
Mm
V (r) = −G (23)
r
Mm
= −G (24)
R+z
Mm
≈ −G + mgz (25)
R
where (26)
M = mass of earth (27)
m = mass of particle (28)
r = distance from center of earth (29)
G = Newton’s gravitational constant (30)
R = radius of earth (31)
z = height of particle above surface of earth (32)
2
g = GM/R . (33)
We may drop the constant term in our discussion, and consider only the
mgz piece, with z ≪ R. We further assume that no angular momen-

5
tum is involved, and treat this as a one dimensional problem. Finally,
assume that the particle is unable to penetrate the earth’s surface.

(a) Make a WKB calculation for the energy spectrum of the particle.
Solution: The potential is

V (z) = mgz. (34)

We need the turning points z1 and z2 :

V (z1 ) = V (z2 ) = E (35)

In this case, since we hit the ground at z = 0, and can go no


further, z1 = 0. The other turning point is at
E
z2 = . (36)
mg

Now we compute the function:


Z z2 (En ) q
f (En ) = 2m [En − V (z)]dz (37)
z1 (En )
Z En /mg q
= 2m(En − mgz)dz (38)
0
1
 
= n+ π. (39)
2
That is,
1 En 1 √
  q Z
n+ π = 2mEn ydy (40)
2 mg 0
s
2 2 3/2
= E . (41)
3g m n
Solving for En , we obtain the estimated bound state energy spec-
trum: !1  2
9π 2 2
3
1 3
En = mg n+ . (42)
8 2

6
(b) If the particle is an atom of atomic weight A ∼ 100, use the result
of part (a) to estimate the particle’s ground state energy (in eV).
Is sunlight likely to move the particle into excited states?
Solution: If A = 100, then m ∼ 100 × 109 eV. Also,
g ∼ 10m/s2 (43)
1
∼ 10m/s2 × × 200MeV-fm × 10−15 m/fm
(3 × 108 m/s)2
∼ 2 × 10−23 eV. (44)
The ground state energy is (n = 0):
!1  2
9π 2 3
1 3
E0 = mg 2 (45)
8 2
−12
∼ 10 eV. (46)
Since photons in sunlight have energies of order eV, they will read-
ily excite such atoms into highly excited states in the gravitational
potential.
(c) Now make a variational calculation for the ground state energy
(i.e., an upper bound thereon). Pick a “sensible” trial wave func-
tion, at least in the sense that it satisfies the right boundary con-
ditions. Compare your result with the ground state level from the
WKB approximation.
Solution: The wave function must vanish at z = 0 and at z = ∞.
A simple trial function which satisfies these boundary conditions
is:
z
ψ(z) = √ e−z/2R , (47)
2R 3

where the variational parameter is R.


We must evaluate the expectation value of the Hamiltonian for
our trial wave function. The kinetic energy part is:
1 d2
!
∞ z z
Z
−z/2R
hT i = √ e − 2
√ e−z/2R dz (48)
0 2R 3 2m dz 2R 3
2
!
1 Z ∞
1 z
= 3
z− e−z/R dz (49)
4mR 0 R 4R
1
= . (50)
8mR2
7
The potential energy part is:
z ∞ z
Z
hV i = e−z/2R mgz √
√ e−z/2R dz (51)
0 2R 3 2R3
= 3mgR. (52)
1
Thus, we wish to minimize the quantity 3mgR+ 8mR 2 as a function

of R. The minimum occurs at


!1
1 3
R= . (53)
12m2 g
Thus, the variational bound on the ground state energy is
" 5 #1/3
3
E0 ≤ mg 2 . (54)
2
We note that this bound is slightly larger than the WKB estimate:
" 5 #1/3 !1/3 1/3
3 9π 2 27

2
mg / mg 2 = . (55)
2 32 π2

4. We discussed the method of stationary phase in section ??. Recall that


the problem it addresses is to evaluate integrals of the form:
Z ∞
I(ǫ) = f (x)eiθ(x)/ǫ dx, (56)
−∞

where f and θ are real, and ǫ > 0. We showed that, in the situation
where ǫ is very small, and θ has a stationary point at x = x0 , this
integral is approximately:
s
√ iθ(x0 )/ǫ i π4 sign[θ ′′ (x0 )] 2π
I(ǫ) = ǫf (x0 )e e [1 + O(ǫ)] . (57)
|θ′′ (x0 )|
If there is more than one stationary point, then the contributions are
to be summed.
To get a little practice applying this method, evaluate the following
integral for large t:
Z 1 h i
J(t) = cos t(x3 − x) dx. (58)
0

8
Solution: To start to get it into the desired form, write
Z 1
ei(t(x ) dx.
3 −x)
J(t) = ℜ (59)
0

Thus, f (x) = 1, θ(x) = x3 − x, and ǫ = 1/t. The first two derivatives


= 3x2 − 1 and θ′′ (x) = 6x.
are θ′ (x) √ √ The first derivative is zero at
x = ±1/ 3. The zero at x0 = 1/ 3 falls within the range of the
integral, so this is the only stationary point
√ of interest. The value of θ
at the stationary point is θ(x0 ) =√−2/3 3. The second derivative at
the stationary point is θ′′ (x0 ) = 2 3.
Plugging into our stationary phase formula, we get:
s
1 − 2it 2π
J(t) ≈ ℜ √ e 3 3 eiπ/4 √

(60)
t 2 3
s  
π i π − 2t

= ℜ √ e 4 3 3 (61)
t 3
s !
π π 2t
= √ cos − √ . (62)
t 3 4 3 3

5. I suggested in section ?? that you consider the classical correspondence


for the time delay (or advance) of the asymptotic motion due to scatter-
ing on a potential. Let us pursue this here. Consider one-dimensional
motion. A particle of mass m is incident from the left on a potential:

V (x) = −K x ∈ (−∆/2, ∆/2) (63)
0 otherwise.
We wish to solve for the motion for large x at large times.

(a) Let’s do the quantum mechanics calculation first. Suppose that


our momentum space wave function at early time is a gaussian
wave packet:
" #1/2
1 − 21 ( p−q )
2
ψ̂(p) = √ e σ . (64)
2πσ
What is ψ(x, t) for large times and large x? Describe the motion,
relative to what it would be if K = 0.

9
(b) Now do the same problem classically. That is, solve for the motion
at large times and large x. Again, compare the result with what
it would be for K = 0. Contrast with the quantum result.

6. We have solved the Schrödinger equation for the Hydrogen atom with
Hamiltonian:
p2 e2
H0 = − .
2m r
The kinetic energy term is non-relativistic – the actual kinetic energy
will have relativistic corrections.

(a) Obtain an expression for the next order relativistic (kinetic en-
ergy) correction to the energy spectrum of hydrogen. It is con-
venient to avoid taking multiple derivatives by using the unper-
turbed Schrödinger equation to eliminate them. Thus, write your
expression in terms of the unperturbed energies and expectation
2 2
values of er and ( er )2 . Do not actually do the integration over
r here, but reduce the problem to such integrals. Make sure you
understand all of your steps.
Solution: The relativistic kinetic energy is
q
T = p2 + m2 − m (65)
q 
= m 1 + (p/m)2 − 1 (66)
1 1
 
= m (p/m)2 − (p/m)4 + O((p/m)6) . (67)
2 8
Thus, the next order relativistic correction to H0 is

1 p4
Hr = − . (68)
8 m3
Following the hint, notice that p4 = 4m2 (H0 − V )2 , where V =
−e2 /r. Thus, the perturbation Hamiltonian may be written
1 1  2 
Hr = − (H0 − V )2 = − H0 − H0 V − V H0 + V 2 . (69)
2m 2m
To determine how the energy levels change in first order pertur-
bation theory, we take the expectation value of Hr with respect

10
to the unperturbed stationary state wave functions:
Er = hψnℓm |Hr |ψnℓm i (70)
1
= − hψnℓm |H02 − H0 V − V H0 + V 2 |ψnℓm i (71)
2m 
1 2 1 4 1

2
= − E + 2En e h i + e h 2 i (72)
2m n r r
1 1 4 1
 
2 2
= − E + 2En e h i + e h 2 i . (73)
2m n r r
(b) Now apply your formula to obtain the first-order relativistic ki-
netic energy correction to the ground state energy of hydrogen.
Express your answer as a multiple of the unperturbed ground state
energy, and also calculate the size of the correction in eV.
Solution: The ground state wave function of hydrogen is:
2
ψ100 (x) = q e−r/a0 (74)
3
4πa0
As there is no dependence anywhere on the angular coordinates,
we are interested in the radial wave function
2
R10 (r) = q e−r/a0 . (75)
a30
Let us evaluate the integral:
Z ∞
Ik ≡ xk e−x dx, k≥0 (76)
0
dk ∞
Z
= lim (−)k e−ax dx (77)
a→1 dxk 0
= k!. (78)
Thus, in the hydrogen atom ground state:
1 4 ∞ 1 −2r/a0 2
Z
h ki = 3 e r dr (79)
r a0 0 r k
1 2 k
 
= (2 − k)!, k ≤ 2 (80)
2 a0
1 1
h i = (81)
r a0
1 2
h 2i = 2 (82)
r a0

11
Thus, the first-order relativistic energy correction to the ground
state enrgy of hydrogen is (noting that E0 = − 12 mα2 is the un-
perturbed ground state energy, and that a0 = 1/mα):
!
1 1 2
Er = − E02 + 2E0 e2 + e4 2 (83)
2m a0 a0
1 1 2 4
 
= − m α − m2 α4 + 2m2 α4 (84)
2m 4
5 2 1 5
 
= − α mα2 = α2 E0 (85)
4 2 4
= 0.91 meV. (86)

7. Let us consider an example of the use of degenerate stationary state


perturbation theory. Thus, let us take the hydrogen atom, with unper-
P2
turbed Hamiltonian H0 = 2m − αr , and consider the effect of putting this
atom in a uniform external electric field: E = Eêz . We are interested
in calculating, to first order in perturbation theory, the shifts in the
n = 2 energy levels. Note that the n = 2 level is four-fold degenerate,
corresponding to the eigenstates: |2S0 i, |2P1i, |2P0 i, |2P−1i, neglecting
spins.

(a) Write down the perturbing potential, V . [Note that we need only
consider the electron’s coordinates, relative to the nucleus – why?]
Calculate the commutator [V, Lz ], and hence determine the matrix
elements of V between states with different eigenvalues of Lz .
Solution: We are interested in computing the shift in the n = 2
energy levels (note that n is the principal quantum number here).
These energy levels are computed with the center-of-mass motion
separated out. Since the hydrogen atom is neutral, there is no
effect on the center-of-mass motion of turning on this electric field.
Hence, we only need consider the relative motion between the
electron and the nucleus. Effectively, we have an electric dipole
interacting with the electric field. The perturbing potential is

V (x) = −eEz = −eEr cos θ. (87)

The commutator [V, Lz ] = −eE[z, xpy − ypx ] = 0. Hence,

0 = hℓ′ m′ |[V, Lz ]|ℓmi = −eE(m − m′ )hℓ′ m′ |z|ℓmi. (88)

12
Thus, hℓ′ m′ |V |ℓmi = δmm′ hℓ′ m|V |ℓmi, that is, the matrix element
is zero unless m = m′ .
(b) You should have found a “selection rule” which simplifies the prob-
lem. What is the degeneracy that needs to be addressed in the
problem now that you have made this calculation?
Solution: The degeneracy that concerns us is the one between
states with like eigenvalues of Lz . These are the states |2S0 i and
|2P0 i.
(c) Using the invariance of the hydrogen atom Hamiltonian under
parity, write down the remaining matrix elements of V which need
to be determined, and compute their values.
Solution: Because V = eEr cos θ is odd under parity, the ex-
pectation value of V is zero between states of like parity. As the
unperturbed Hamiltonian commutes with parity, its eigenstates
may be expressed as eigenstates of parity. In particular, the P
states have odd parity, and the S states have even parity. Hence
the only (potentially) non-zero matrix elements of V in the four-
dimensional subspace we are considering here are h2S0 |V |2P0 i and
h2P0 |V |2S0 i. Since our wave functions are real (by convention),
these two matrix elements are equal.
The hydrogenic wave functions we need are (from solutions to
problem 40):
1 1 1
ψ200 = √ √ 3/2 (2 − r/a0 )e−r/2a0 (89)
4π 2 2 a0
s
3 1 1 r
ψ210 = cos θ √ 3/2 2 e−r/2a0 . (90)
4π 4 6 a0 a0

Recall also, from the solution to exercise 4:


Z ∞
Ik ≡ xk e−x dx, k≥0 (91)
0
= k!. (92)

Thus, the desired matrix element is:


√ Z
3 2π
Z 1
h2P0 |V |2S0 i = −eE dφ cos2 θd cos θ
4π 0 −1

13
1 1 Z∞ 2 r
√ 3 r drr(2 − r/a0 ) e−r/a0 (93)
8 3 a0 0 a0
a0 Z ∞ 4
= −eE x (2 − x)e−x dx (94)
3·8 0
a0
= −eE (2 · 4! − 5!) (95)
3·8
= 3eEa0 . (96)
The perturbing Hamiltonian may thus be written, in the |2S0 i, |2P1i, |2P0 i, |2P−1i
basis:
0 0 1 0
 
0 0 0 0
V = 3eEa0  . (97)
 
1 0 0 0
0 0 0 0
(d) Now complete your degenerate perturbation theory calculation to
determine the splitting of the states in the applied electric field.
Calculate numerical splittings (in eV) for an applied field of 100
kV/cm. Also, estimate the “typical” electric field felt by the elec-
tron, due to the nucleus, in a hydrogen atom. Was the use of
perturbation theory reasonable for this problem?
Solution: The eigenstates of the perturbing Hamiltonian are:
1
√ (|2S0i + |2P0 i) (98)
2
1
√ (|2S0 i − |2P0 i) (99)
2
|2P1 i (100)
|2P1 i, (101)
with eigenvalues 3eEa0 , −3eEa0 , 0, 0, respectively. In a 100
kV/cm field,
3eEa0 ∼ 3 × 100 keV/cm0.5 × 10−8 cm (102)
= 1.5 × 10−3 eV. (103)
The “typical” electric field felt by the electron in the field of the
proton is:
e2 /r 2 × −13.6 eV
Ep ∼ = (104)
er e × 0.5 × 10−8 cm
9
∼ 5 × 10 kV/cm. (105)

14
The use of perturbation theory for this problem appears justified.

8. It may happen that we encounter a situation where the eigenvalues of


H0 , call them εn and εm , are nearly, but not quite equal. In this case,
we cannot use degenerate perturbation theory, and ordinary perturba-
tion theory looks unreliable. Let us try to deal with such a situation:
Suppose the two eigenstates |ni and |mi of H0 have nearly the same en-
ergy (and all other eigenstates don’t suffer this disease, for simplicity).
Let H = H0 + V , and write
X
V = |iihi|V |jihj| (106)
i,j
H0 |ii = εi |ii, (107)
where

hi|ji = δij· (108)


Let

V = V1 + V2 , (109)
with

V1 ≡ |mihm|V |mihm| + |nihn|V |nihn| + (110)


+|mihm|V |nihn| + |nihn|V |mihm| (111)

and V2 is everything else.


If we can solve exactly the problem with H1 = H0 + V1 , then the trou-
blesome 1/(εn − εm ) terms are avoided by the exact treatment, and we
may treat V2 as a perturbation in ordinary perturbation theory (since
hi|V2 |ji = 0 for i, j = n, m). All states |ii, i 6= n, m, are eigenstates
of H1 , since V1 |ii = 0 in this case. However, |ni and |mi are not in
general eigenstates of H1 .

(a) Solve exactly for the eigenstates and eigenvalues of H1 , in the


subspace spanned by |ni, |mi. Express your answer in terms of

εn , εm , hm|V |ni, hn|V |ni, hm|V |mi.

15
(You may also use the shorthand
(1)
En,m = εn,m + hn, m|V |n, mi
if you find it convenient.)
Solution: A vector |ii in our restricted subspace is of the form:
|ii = α|mi + β|ni, (112)
where |mi and |ni are eigenstates of H0 with eigenvalues εm and
εn , respectively. Schrödinger’s equation with Hamiltonian H1 =
H0 + V1 is:
εm + hm|V |mi hm|V |ni α α
    
=E . (113)
hn|V |mi εn + hn|V |ni β β
We find eigenvalues E by taking the determinant:
εm + hm|V |mi − E hm|V |ni


= 0. (114)
hn|V1 |mi εn + hn|V |ni − E
Letting Vij ≡ hi|V |ji = Vji∗ , we find eigenvalues E± :
1 (1)
 q 
(1) (1)
E± = E + En(1) ± (Em − En )2 + 4|Vmn |2 . (115)
2 m
The corresponding eigenvectors are determined by:
(1)
Em α± + Vmn β± = E± α± (116)
(1)
En β± + Vnm α± = E± β± (117)
Hence,
(1)
E± − Em
β± = α± . (118)
Vmn
Imposing the normalization |α± |2 + |β± |2 = 1, and choosing α±
real and positive, we have:
v
u (1)
u (E± − Em )2
α± = 1/ 1 +
t
(119)
|Vmn |2
v !2
Vmn 2
Vmn
u
u
β± = 1/t + (1)
(120)
|Vmn | (E± − Em )2
The eigenvectors are thus,
|±i = α± |mi + β± |ni. (121)

16
(b) As an application, consider an electron in a weak one-dimensional
periodic potential (“lattice”) V (x) = V (x + d). Assume the lat-
tice has a size L = Nd, and that we the have periodic boundary
condition on our wave functions: ψ(x) = ψ(x + L). With this
boundary condition, the unperturbed wave functions are plane
waves, ψp (x) = √1L eipx , where p = 2πn/L, n=integer, and the
 2
p2 2πn 1
unperturbed eigenenergies are εn = 2m
= L 2m
. We expand
the potential in a Fourier series:

ein2πx/d Vn
X
V (x) =
n=−∞

If we label our eigenfunctions by |pi = √1 e2πinp x/L , determine all


L
nonvanishing matrix elements of V :

hq|V |pi

Express your answer in terms of Vn .


Solution:
∞ L 1
Z
ein2πx/d √ e2πi(np −nq )x/L dx (122)
X
hq|V |pi = Vn
n=−∞ 0 L
∞ Z 1
ei2π(np −nq +N n)u du
X
= Vn (123)
n=−∞ 0

X∞
= Vn δN n,(nq −np ) . (124)
n=−∞

Thus, hq|V |pi = 6 0 if and only if (nq − np )/N is an integer, and


V(nq −np )/N 6= 0.
(c) Suppose εnp and εnq are not close to each other ∀nq , given some
np . Calculate the perturbed wave function in ordinary first order
perturbation theory corresponding to unperturbed wave function
ψp (x). Also, calculate the energy to 2nd order. Express your
answer in terms of Vn and the unperturbed energies.
Solution: The first order perturbed wave function is:
X hq|V |pi
Ψp (x) = ψp (x) + |qi . (125)
q6=p ε np − ε nq

17
2π 2
2 2
Now εnp − εnq = mL 2 (np − nq ), and hq|V |pi = 0 unless nq =

np + Nn. Thus, letting |P i = Ψp (x):


X Vn
|P i = |pi + |np + nNi (126)
nq =np +nN ε np − ε nq

X Vn
|P i = |pi + |np + nNi . (127)
n=−∞,6=0
ε np − εnp +nN
Alternatively,

X Vn
|P i = |pi + |np + nNi 2π 2
(128)
n=−∞,6=0 mL2
(n2p − n2q )
2 ∞
md X Vn
= |pi + |np + nNi (129)
2π 2 n=−∞,6=0 (n2p − n2q )/N 2
2 ∞
md X Vn
= |pi − |np + nNi (130)
2π 2 n=−∞,6=0
n(n + 2np /N)
2 ∞
md X Vn
= |pi − |np + nNi (131)
2π 2 n=−∞,6=0 n(n + 2np /N)
 

md2 Vn
= |pi 1 − e−i2πnp x/L 2
X
exp i2π(np + nN)x/L
2π n=−∞,6=0 n(n + 2np /N)
 

md2 X Vn
= |pi 1 − exp i2πnx/d . (132)
2π 2 n=−∞,6=0 n(n + 2np /N)

The first order energy correction is


Ep1 = hp|V |pi = V0 . (133)
The second order energy correction is
X |hq|V |pi|2 X |Vn |2
Ep2 = = (134)
q6=p εnp − εnq q6=p εnp − εnq

X md2 |Vn |2
= − 2 n(n + 2n /N)
. (135)
n=−∞,6=0
2π p

Thus, the energy to second order is:



md2 X |Vn |2
Ep = εnp + V0 − (136)
2π 2 n=−∞,6=0 n(n + 2np /N)

18
2π 2 n2p md2 ∞
X |Vn |2
= + V 0 − (137)
mL2 2π 2 n=−∞,6=0
n(n + 2np /N)

(d) What is the condition on np (and hence on p) so that |pi will be


nearly degenerate in energy with another eigenstate of H0 ?
Solution: From part (c), we see that the coefficient in the ex-
panison for |P i blows up when n + 2np /N = 0, i.e., at
Nn
np = − , n = ±1, ±2, . . . . (138)
2
When np takes on such a value, p = 2πnp /L = −πn/d. This is
degenerate with q = πn/d = p + 2πn/d.
(e) Assume that the condition in (d) exists, and use part (a) to solve
this “almost degenerate” case for the eigenenergies. Complete the
graph in Fig. 1 for higher values of |p|.

Ep

|V0 |

p
- π /d 0 π /d

Figure 1: Energy versus momentum for the one-dimensional lattice problem


(6).

Solution: Let |pi be nearly degenerate with |qi:


(n + ǫ)π 2πn (n − ǫ)π
p=− , q =p+ = , (139)
d d d
19
where 0 < |ǫ| ≪ 1. The nearly degenerate energies are:

π2 π2
ε np = (n + ǫ)2 , ε nq = (n − ǫ)2 . (140)
2md2 2md2
The matrix elements we need are:

hp|V |pi = hq|V |qi = V0 , (141)


hp|V |qi = hq|V |pi = Vn . (142)

The energies, using part (a), are


1h q i
E± = εnp + V0 + εnq + V0 ± (εnp − εnq )2 + 4|Vn |2 (143)
2 v
!2
π2 u π2n
u
2 2
= 2
(n + ǫ ) + V 0 ± t
ǫ2 + |Vn |2 . (144)
2md md2

The difference in energy between these two values is


v
!2
π2n
u
u
E+ − E1 = 2 t
ǫ2 + |Vn |2 (145)
md2
!2
1 π2n
≈ 2|Vn | + ǫ2 . (146)
|Vn | md2

9. When we calculated the density of states for a free particle, we used a


“box” of length L (in one dimension), and imposed periodic boundary
conditions to ensure no net flux of particles into or out of the box.
We have in mind that we can eventually let L → ∞, and are really
interested in quantities per unit length (or volume). Let us justify
more carefully the use of periodic boundary conditions, i.e., we wish to
convince ourselves that the intuitive rationale given above is correct.
To do this, consider a free particle in a one-dimensional “box” from
−L/2 to L/2. Remembering that the Hilbert space of allowed states is
a linear space, show that the periodic boundary condition:

ψ(−L/2) = ψ(L/2), (147)


ψ ′ (−L/2) = ψ ′ (L/2) (148)

20
gives acceptable wave functions. “Acceptable” here means that the
probability to find a particle in the box must be constant. Are there
other acceptable choices?
Solution: The Schrödinger equation for a free particle is
1 2
−i∂t ψ(x, t) = − ∂ ψ(x, t). (149)
2m x
We suppose that an “acceptable” wave function is one which has a
constant probability to be in the “box” (−L/2, L/2):
d L/2
Z
|ψ(x, t)|2 dx = 0. (150)
dt −L/2
It is readily verified that the function
2π 2 2π
φ(x, t) = ei mL2 t sin x (151)
L
has the desired property.
If we admit φ(x, t) as an acceptable solution, and if ψ(x, t) is any other
acceptable solution, then φ + ψ must be acceptable, since any linear
combination of acceptable solutions must be acceptable. Hence, we
must have:
d L/2
Z
|ψ(x, t)|2 dx = 0; (152)
dt −L/2
d L/2
Z
|φ(x, t)|2 dx = 0; (153)
dt −L/2
d L/2
Z
|ψ(x, t) + φ(x, t)|2 dx = 0. (154)
dt −L/2
Then we may write (assuming Eqns 152 and 153):
d Z L/2
0 = [ψ(x, t)φ∗ (x, t) + ψ ∗ (x, t)φ(x, t)] dx (155)
dt −L/2
Z L/2
= ∂t [ψ(x, t)φ∗ (x, t) + ψ ∗ (x, t)φ(x, t)] dx (156)
−L/2
i Z L/2 h 2  ∗       i
= ∂x ψ φ − ψ ∂x2 φ + ψ ∗ ∂x2 φ − ∂x2 ψ ∗ φ dx
(157)
2m −L/2
Z L/2
= ∂x [(∂x ψ) φ∗ − ψ (∂x φ) + ψ ∗ (∂x φ) − (∂x ψ ∗ ) φ] dx (158)
−L/2
L/2
= [(∂x ψ) φ∗ − ψ (∂x φ) + ψ ∗ (∂x φ) − (∂x ψ ∗ ) φ]−L/2 . (159)

21
But φ(±L/2, t) = 0, so
L/2
0 = [−ψ(∂x φ∗ ) + ψ ∗ (∂x φ)]−L/2 . (160)
Further, since
2π i 2π22 t
∂x φ(±L/2, t) = − e mL , (161)
L
we obtain
2π 2 2π 2 2π 2 2π 2
0 = ψ(L/2, t)e−i mL2 t −ψ(−L/2, t)e−i mL2 t +ψ ∗ (L/2, t)ei mL2 t −ψ ∗ (−L/2, t)ei mL2 t .
(162)

This must be true for all times; also if ψ is acceptable, then e ψ must
be acceptable, for real θ. Hence, ψ is acceptable if and only if Eqn. 152
holds, and:
ψ(L/2, t) = ψ(−L/2, t). (163)
2π 2
We note that the function ei mL2 t cos 2π
L
x satisfies these criteria. Thus,
we could also have picked

2π 2
φ(x, t) = ei mL2 t cos
x (164)
L
as an acceptable solution. Then the same argument reveals that any
other acceptable solution ψ must satisfy the boundary condition:
∂x ψ(L/2, t) = ∂x ψ(−L/2, t). (165)
2 2π 2 2 2
 
in t n 2π
We finally remark that the set of functions e mL2 sin 2πn
L
x, ei mL2 t cos 2πn
L
x; n = 0, 1, . . .
is a complete set of functions with the required boundary conditions.
10. See if you can generalize the result for the first Born approximation:
dσ m2
= |V̂ (p′ − p)|2 . (166)
dΩ′ (2π)2
to the case where the scattered particle (mass mf ) may have a different
mass than the incident particle (mass mi ).
Solution:

dσ mi mf pf

= 2
|V̂ (pf − pi )|2 . (167)
dΩ (2π) pi

22
11. We consider the potential (called the “Yukawa potential”):
Ke−µr
V (x) = , r = |x|,
r
with real parameters K and µ > 0. The parameter K can be regarded
as the “strength” of the potential (“interaction”), and µ1 is effectively
the “range” of distance over which the potential is important. µ itself
has units of mass – note that as µ → 0 we obtain the Coulomb potential:
µ can be thought of as the mass of an “exchanged particle” which
mediates the force. In electromagnetism, this is the photon, hence
µ → mγ = 0

(a) Find a condition on K and µ which guarantees that there are at


least n bound states in this potential. You will likely fashion and
use some kind of “comparison” theorem in arriving at your result.
You should give at least a “heuristically convincing” argument, if
you don’t actually prove it.
Solution: We assume K < 0, so that the potential is attractive.
Consider
e−µr 1
!
D(r) ≡ K − . (168)
r r
For r > 0, 0 < e−µr ≤ 1, hence 0 ≤ D(r). Also
dD(r) Kh i
= − 2 (1 + µr)e−µr − 1 < 0. (169)
dr r
Hence, D(r) decreases monotonically for r > 0, with a maximum
at r = 0:
lim D(r) = −Kµ. (170)
r→0

Thus, 0 ≤ D(r) ≤ −Kµ.


We may write
e−µr 1
!
K
V (r) = +K − . (171)
r r r
Hence, E ≤ E(hydrogen) + (−Kµ). So, the nth level of V is a
bound state, En < 0, if

En (hydrogen) − Kµ < 0 (172)

23
mK 2
− − Kµ < 0 (173)
2n2
µ m
− < 2. (174)
K 2n

(b) Using the Born approximation for the differential cross section
that we developed in our discussion of time-dependent perturba-

tion theory, calculate the differential cross section, dΩ , for scatter-
ing on this potential. Consider the limit µ → 0 and compare with
the Coulomb differential cross section we obtained in the notes.
Solution:

4πK
V̂ (p) = . (175)
µ2 + p2
dσ m2 ′ 2 4m2 K 2
= |V (p − p)| = 2 . (176)
dΩ (2π)2

µ2 + 4p2 sin2 2θ

(c) Integrate your differential cross section over all solid angles to
obtain the “total cross section”. Again, consider the limit µ → 0.
Hence, what is the total cross section for scattering on a Coulomb
potential?
Solution:

16πm2 K 2
σT = . (177)
µ2 (µ2 + 4p2 )

24

Anda mungkin juga menyukai