Phys4011 W13 PDF
Phys4011 W13 PDF
Phys4011 W13 PDF
Lecture Notes
Tom Kirchner1
April 7, 2013
1 [email protected]
Contents
5 Molecules 70
5.1 The adiabatic (Born-Oppenheimer) approximation . . . . . . . 71
5.2 Nuclear wave equation: rotations and vibrations . . . . . . . . 74
5.3 The hydrogen molecular ion H+ 2 . . . . . . . . . . . . . . . . . 77
1
Chapter 1
ĤΨ = EΨ (1.1)
for the two-body problem consisting of a nucleus (n) and an electron (e).
The Hamiltonian reads
p̂2n p̂2e Ze2
Ĥ = + − (1.2)
2mn 2me 4πǫ0 |re − rn |
with
mn = N 1836me ; me ≈ 9.1 × 10−31 kg
and N being the number of nucleons (N = 1 for the hydrogen atom itself,
where the nucleus is a single proton).
The first step is to separate this two-body problem into two effective
one-body problems.
(re , pe , rn , pn ) −→ (R, P, r, p)
2
3
definitions : M = me + mn ≈ mn
me mn
µ = ≈ me
me + mn
R = mn rnM+me re ≈ rn
center − of − mass
motion
P = pe + pn = M Ṙ ≈ pn
r = re − rn
relative
motion
p = µṙ = mn peM
−me pn
≈ pe
• QM transformation analogously
(r̂e , p̂e , r̂n , p̂n ) −→ (R̂, P̂, r̂, p̂) (1.3)
insertion into Eq. (1.2) yields
l̂ = r̂ × p̂. (1.10)
֒→ they have a common set of eigenstates. The eigenstates of l̂2 , ˆlz are the
spherical harmonics Ylm .
insertion into Eq. (1.7) for Hamiltonian (1.9) yields the radial SE
( )
p̂2r ~2 l(l + 1)
+ + V (r) − Erel Rl (r) = 0 (1.13)
2µ 2µr 2
~2
with p̂2r = − ∂r (r 2 ∂r )
r2
l̂2
and operator identity p̂2 = p̂2r + 2
r
(can be proven in spherical coordinates in coordinate space).
(1.13)
h l(l + 1) i
֒→ yl′′(r) + ǫ − U(r) − yl (r) = 0 (1.15)
r2
~2 ~2
Erel = ǫ, V (r) = U(r)
2µ 2µ
The radial Eq. (1.15) is very similar to the one-dimensional SE. There are,
however, two important differences. First, the total (effective) potential con-
sists of two parts
l(l + 1) r→∞
Ulef f (r) = U(r) + −→ 0
r2
տ ′′ angular momentum barrier′′
(cf. classical central-field problem) Second, the boundary conditions are dif-
ferent.
a) Boundary conditions
• r −→ 0
|ϕrel (r)|2 = |Rl (r)|2|Ylm (θ, ϕ)|2 < ∞
in particular for r = 0
֒→ ’regularity condition’
yl (0) = 0 (1.16)
• r −→ ∞
1. Erel < 0 (bound spectrum)
Z Z ∞ Z
2 3
|ϕrel (r)| d r = 2 2
r Rl (r) dr |Ylm (θ, ϕ)|2 dΩ
Z0 ∞
= yl2 (r) dr < ∞
0
note: for Erel > 0 the solution leads to Rutherford’s scattering formula
(which is identical in classical mechanics and QM)
b) Bound-state solutions
definition : κ2 = −ǫ > 0
4πǫ0 ~2
a = ≈ 0.53 · 10−10 m
µe2
for µ ≡ me , a ≡ a0 is the ′ Bohr radius′
transformation: x = 2κr
d2 1 d2
֒→ =
dx2 4κ2 dr 2
!
d2 l(l + 1) λ 1
֒→ − + − yl (x) = 0 (1.18)
dx2 x2 x 4
asymptotic solutions:
1. x −→ ∞ !
d2 1
֒→ − yl (x) = 0
dx2 4
x x
֒→ yl (x) = Ae− 2 + Be 2
because of yl (x → ∞) = 0 ֒→ B = 0
7
2. x −→ 0 !
d2 l(l + 1)
֒→ − yl (x) = 0
dx2 x2
A
֒→ yl (x) = + Bxl+1
xl
because of yl (0) = 0 ֒→ A = 0 .
Insertion into Eq. (1.18) yields a new differential equation for vl (x):
( )
d2 d
֒→ x 2 + (2l + 2 − x) − (l + 1 − λ) vl (x) = 0. (1.20)
dx dx
More specifically:
nr = n − l − 1 ≥ 0
vl (x) = L2l+1
n−l−1 (x) ,
⇐⇒ n − 1 ≥ l
Z
with n ≡ λn = , n = 1, 2, ... (1.21)
κn a
The detailed treatment shows that the integrability of the solutions requires
Z
λ = κa to be positive, integer numbers −→ quantization of κ (i.e. quantiza-
tion of the energy)1
(n − l − 1)! l+ 1 l+2 √
ϕrel (r) ≡ ϕnlm (r) = 3 2 2 κn a
(n + l)!
× r l e−κn r L2l+1
n−l−1 (2κn r)Ylm (θ, ϕ) (1.22)
n≥0
≡ Rnl (r)Ylm (θ, ϕ) l ≤n−1
−l ≤ m ≤ l
~2 Z 2
Erel ≡ En = − n = 1, 2, ... (1.23)
2µa2 n2
Z2
≈ −13.6 eV 2 . (1.24)
n
with R = 13.6 eV
9
E s p d
0
n = 3
3 s 3 p -1 3 p 0 3 p 1 3 d -2 3 d -1 3 d 0 3 d 1 3 d 2
n = 2
2 s 2 p -1 2 p 0 2 p 1
n = 1
1 s
given n l = 0, 1, ..., n − 1
given l m = −l, ..., l
n−1
X
֒→ (2l + 1) = n2
l=0
In QM, the wave functions themselves are (usually) not observable, but
their absolute squares are
2
= ynl (r) (1.26)
ρnl (r)dr is the probability to find the electron in the interval [r, r + dr]
Momentum space representation
So far we have worked in coordinate space, in which states are wave func-
tions ϕ(r). It is also possible — and insightful — to look at the problem
in another, e.g., the momentum space representation, which is connected to
the coordinate space representation by a (three-dimensional) Fourier trans-
formation. Using the Dirac notation one can obtain momentum space wave
functions by considering
Z
ϕnlm (p) = hp|ϕnlmi = hp|rihr|ϕnlmid3r
Z
1 i
= 3 e− ~ p·r ϕnlm (r)d3 r. (1.27)
[2π~] 2
To work out the three-dimensional Fourier transform one uses the expan-
sion of a plane wave in spherical coordinates
L
∞ X
X
e−ik·r
= 4π (−i)L jL (kr)YLM (Ωk )YLM
∗
(Ωr ) (1.28)
L=0 M =−L
11
Figure 1.2: Radial hydrogen 1s, 2s, 2p wave functions (blue) and probability
densities (red) in coordinate space.
do not approach zero for r → 0). To compensate for the stronger binding
energy of the 2s state at small r the 2p state has to have its only maximum
at smaller r compared to the second maximum of the 2s state.
Figure 1.3: Radial hydrogen 3s, 3p, 3d wave functions (blue) and probability
densities (red) in coordinate space.
Figure 1.4: Radial hydrogen 1s, 2s, 2p wave functions (blue) and probability
densities (red) in momentum space.
Figure 1.5: Radial hydrogen 3s, 3p, 3d wave functions (blue) and probability
densities (red) in momentum space.
b) Hydrogen-like ions
We have solved not only the (Schrödinger) hydrogen problem (Z = 1),
but also the bound-state problems of all one-electron atomic ions (e.g.
He+ , Li2+ ,...) for Z = 2, 3, ... Note that En ∝ Z 2 .
c) Exotic systems
... are also solved
In these cases one has to take care of the different masses compared to
the hydrogen problem. Note that En ∝ µ = mm11+m m2
2
.
d) Corrections
The spectrum determined by Eq. (1.23) is the exact solution of the
Schrödinger-Coulomb problem, but not exactly what one sees experi-
mentally. The reason is that the Schrödinger equation is not the ulti-
mate answer, e.g., it has to be modified to meet the requirements of
the theory of special relativity. Therefore, corrections show up, which
lead to a (partial) lifting of the degeneracy. This will be discussed later
on.
e) Atomic units
So far, we have used SI units (as we are supposed to). In atomic and
molecular physics another set of units is more convenient and widely
used: atomic units. The starting point for their definition is the Hamil-
tonian
~2 2 e2
ĤSI = − ∇ − (1.31)
2me r 4πǫ0 r
(i.e., the one of Eq. (1.7) for Z = 1 and µ → me ). Four constants show
up in this Hamiltonian — way too many — and so they are all made
to disappear!
Recipe
Consequences
• Ĥa.u. = − 12 ∇2r − 1
r
15
4πǫ0 ~2
a0 = = 0.53 · 10−10 m = 1a.u.
me e2
~2
E1s = − = −13.6 eV = −0.5 a.u. = −0.5 hartree = −1 Rydberg
2me a20
e2 1
α= ≈ .
4πǫ0 ~c 137
In atomic units we have α = 1/c, i.e. c ≈ 137 a.u. Thus, one
atomic unit of velocity corresponds to 2.2 · 106 m/s. This is also
obtained by using v0 = a0 /t0 .
Chapter 2
Taking a look at the total potential (Coulomb + Stark) one finds that
tunnelling is possible — i.e. in a strict sense the Hamiltonian (2.5) does not
support stationary states. Eventually, a bound electron will tunnel through
the barrier and escape from the atom. In practice, however, the Stark po-
tential is weak compared to the Coulomb potential and the tunnel effect is
unimportant unless one studies highly-excited states. This is why we can
apply stationary perturbation theory (PT).
decompose
Ĥ = Ĥ0 + Ŵ (2.7)
and assume that the eigenvalue problem of Ĥ0 is known
We seek solutions of. Eq. (2.6) in terms of a Taylor (like) expansion based on
the (nondegenerate) eigenvalues and eigenstates of the ’unperturbed problem’
Eq. (2.8). Therefore, we require that the ’perturbation’ Ŵ be small. Let’s
introduce a smallness parameter λ:
d
Ĥ0 + λŵ − Eα (λ) |ϕα (λ)i = 0
dλ
⇐⇒ Ĥ0 + λŵ − Eα (λ) |ϕ′α (λ)i + ŵ − Eα′ (λ) |ϕα (λ)i = 0
dEα
(Eα′ = etc.)
dλ
֒→ hϕβ (λ)|Ĥ(λ) − Eα (λ)|ϕ′α (λ)i + hϕβ (λ)|ŵ − Eα′ (λ)|ϕα (λ)i = 0
i) α = β
=⇒ Eα′ (λ) = hϕα (λ)|ŵ|ϕα (λ)i (2.13)
ii) α 6= β
d
proof : hϕα (λ)|ϕα (λ)i = hϕ′α (λ)|ϕα (λ)i + hϕα (λ)|ϕ′α (λ)i
dλ | {z }
=1
= 2hϕα (λ)|ϕ′α(λ)i = 0
X
֒→ |ϕ′α (λ)i = |ϕβ (λ)ihϕβ (λ)|ϕ′α (λ)i
β
X hϕβ (λ)|ŵ|ϕα (λ)i
= |ϕβ (λ)i (2.15)
β6=α
Eα (λ) − Eβ (λ)
19
Eqs. (2.17), (2.18) are the standard expressions for the lowest-order correc-
tions — the glorious result of this section!
Remarks:
(0) (0)
1. Derivation and result are valid only if Eα 6= Eβ (i.e. no degeneracies)
2. Convergence of perturbation series?
This cannot be answered in general. In some cases, perturbation ex-
pansions do converge, in some they do not, and in some other cases the
perturbation series turns out to be a so-called semi-convergent (asymp-
totic) series.
Consistency criterion for convergence (cf. Eq. (2.18))
hϕ0 |Ŵ |ϕ0 i
β α
(0) (0) ≪1, (for α 6= β)
Eα − E
β
20
3. In practice, ’exact’ calculations beyond 1st order in the energy are often
not feasible due to (infinite) sums over all basis states (cf. Eq. (2.17)).
1
ϕ01s (r) = √ e−r
π
(0)
E1s = −0.5 a.u.
W = Fz
(with x = cos θ)
The 2nd -order energy correction
(2)
X |hϕ01s |F z|ϕ0β i|2
∆E1s = (0) (0)
β6=1s E1s − Eβ
is hard to calculate due to the infinite sum (which actually also involves
an integral over the continuum states). Let us content ourselves with an
estimate.
(0) (0) (2)
Note that E1s − Eβ < 0, i.e., ∆E1s < 0.
21
Consider
(2)
X |hϕ01s |z|ϕ0β i|2 F2 X
2
|∆E1s | = F (0) (0)
< (0) (0)
hϕ01s |z|ϕ0β ihϕ0β |z|ϕ01s i
β6=1s Eβ − E1s En=2 − E1s β6=1s
!
F2 X
= (0) (0)
hϕ01s |z |ϕ0β ihϕ0β |z|ϕ01s i − hϕ01s |z|ϕ01s ihϕ01s |z|ϕ01s i
En=2 − E1s β
2 2
8F (2.20) 8F
= (hϕ01s |z 2 |ϕ01s i − hϕ01s |z|ϕ01s i2 ) = hϕ01s |z 2 |ϕ01s i
3 3
One finds hϕ01s |z 2 |ϕ01s i = 1 and obtains
(2) 8
|∆E1s | < F 2 quadratic Stark effect (2.21)
3
The exact result (see [Sha], Chap. 17) is
(2) 9
∆E1s = − F 2 (2.22)
4
Interpretation:
Consider a classical charge distribution ρ in an electric field. The associated
potential energy is
Z Z
W = ρ(r)Φ(r)d r = −F ρ(r)zd3 r
3
= −F dz
Summary:
(1)
• ∆E1s = 0 expresses the fact that the unperturbed hydrogen ground
(0)
state has no static dipole moment dz (this is in fact true for any
spherically symmetric charge distribution).
(2)
• ∆E1s 6= 0 expresses the fact that it has a nonzero induced dipole
(1) (1)
moment dz , i.e., a nonzero dipole polarizability αD := dz /F . We
have found αD < 16/3 a.u. (the exact result being αD = 9/2 a.u. )
The unknown states |ϕ̃0αj i are the 0th -order states of the pertubation expan-
sion:
(1)
Eαj = Eα(0) + λEαj + · · · (2.27)
|ϕαj i = |ϕ̃0αj i + λ|ϕ1αj i + · · · (2.28)
23
Note that this expansion is of the same type as the previous Taylor series of
Eqs. (2.11), (2.12) if one identifies
(1) (1) dEα (λ)
∆Eα = λEα = λ etc.
dλ λ=0
Now proceed as follows: Insert (2.27), (2.28) into the Schrödinger equation
(2.25) to obtain
(Ĥ0 + λŵ) |ϕ̃0αj i + λ|ϕ1αj i + · · ·
(1)
= (Eα(0) + λEαj + · · · ) |ϕ̃0αj i + λ|ϕ1αj i + · · · ,
sort this in terms of powers of λ
λ0 : Ĥ0 |ϕ̃0αj i = Eα(0) |ϕ̃0αj i
(1)
λ1 : Ĥ0 |ϕ1αj i + ŵ|ϕ̃0αj i = Eα(0) |ϕ1αj i + Eαj |ϕ̃0αj i,
and project the second equation onto an undisturbed eigenstate of the same
subspace:
(1)
0 = hϕ0αl |Ĥ0 − Eα(0) |ϕ1αj i + hϕ0αl |ŵ − Eαj |ϕ̃0αj i
(1)
⇔ 0 = hϕ0αl |ŵ − Eαj |ϕ̃0αj i
gα
X (1)
⇔ 0 = hϕ0αl |ŵ − Eαj |ϕ0αk iaαkj l = 1, . . . , gα (2.29)
k=1
λ ≡ F (smallness parameter)
ϕnlm (r) = Rnl (r)Ylm (Ω)
r
4π
z = r cos θ = rY10 (Ω)
3
r Z ∞ Z
0 0 4π 3 ∗
֒→ hϕnlm |z|ϕnl′ m′ i = r Rnl (r)Rnl′ (r)dr Ylm (Ω)Y10 (Ω)Yl′ m′ (Ω)dΩ
3 0
The angular integral is a special case of a more general integral over three (ar-
bitrary) spherical harmonics, the result of which is known (”Wigner-Eckart
theorem”):
r Z
4π ∗
Ylm (Ω)YLM (Ω)Yl′ m′ (Ω)dΩ
2L + 1
m
p
′ l L l′ l L l′
= (−1) (2l + 1)(2l + 1) (2.30)
−m M m′ 0 0 0
with
j1 j2 j3
∈R ”Wigner’s 3j-symbol”
m1 m2 m3
The 3j-symbols are closely related to the Clebsch-Gordan coefficients. We
don’t need to know much detail here other than that they fulfill certain
selection rules, i.e., they are zero in many cases:
25
j1 j2 j3
6= 0 iff m1 + m2 + m3 = 0 ∧ |j1 − j2 | ≤ j3 ≤ j1 + j2
m1 m2 m3
j1 j2 j3
6 = 0 iff j1 + j2 + j3 = even ∧ |j1 − j2 | ≤ j3 ≤ j1 + j2
0 0 0
m = m′ and ∆l = l − l′ = ±1 (2.31)
• matrix elements:
due to the dipole selection rules (2.31) there is only one nonzero matrix
element we need to calculate:
Z ∞ Z 1 Z 2π
0 0 1 4 −r 2
w12 = hϕ2s |r cos θ|ϕ2p0 i = (2 − r)r e dr cos θ d cos θ dϕ
32π 0 −1 0
Z ∞ Z ∞
1 4 −r 5 −r 1
= 2 r e dr − r e dr = (2 × 4! − 5!) = −3 (a.u.)
24 0 0 24
= w21
→ perturbation matrix:
0 w12 0 0
w12 0 0 0
w=
0
0 0 0
0 0 0 0
26
• secular equation:
−E (1) w12 0 0
w12 −E (1) 0 0
det
0 (1)
= 0 (2.32)
0 −E 0
0 0 0 −E (1)
⇔ (E (1) )4 − (E (1) )2 w12
2
= 0
(1)
֒→ E = {0, 0, w12, −w12 }
• mixing coefficients
(1)
(i) E1,2 = 0:
(1,2)
0 w12 0 0 a2s 0
(1,2)
w12 0 0 0
a2p0 0
=
0 a(1,2)
0
0 0 2p−1 0
0 0 0 0 (1,2)
a2p+1 0
(1,2) (1,2)
⇔ a2s = a2p0 = 0
(1,2) (1,2)
֒→ |ϕ̃0E (1) i = a2p−1 |ϕ02p−1 i + a2p+1 |ϕ02p+1 i
1,2
(1)
(ii) E3 = w12 :
(3)
−w12 w12 0 0 a2s 0
(3)
w12 −w12 0 0
a2p0 0
=
0 a(3)
0
0 0 2p−1 0
0 0 0 0 (3)
a2p+1 0
(3) (3) (3) (3)
⇔ a2s = a2p0 , a2p−1 = a2p+1 = 0
27
(1)
(iii) E4 = −w12 :
(4)
w12 w12 0 0 a2s 0
(4)
w12 w12 0 0
a2p0 0
=
0 0 a(4)
0
0 2p−1 0
0 0 0 0 (4)
a2p+1 0
(4) (4) (4) (4)
⇔ a2s = −a2p0 , a2p−1 = a2p+1 = 0
Summary
1. The (weak) electric field results in a splitting of the energy level — the
degeneracy is (partly) lifted. The energy shifts are linear in the electric
field strength:
(1)
∆E1,2 = 0
(1)
∆E3 = λw12 = −3F
(1)
∆E4 = −λw12 = 3F
2. Note that the ’original’ L-shell states have no static dipole moment
since
hϕ0nlm |z|ϕ0nlm i = 0
3. The Stark states (2.33), (2.34) do have nonzero static dipole moments
(calculate them!). Check the Maple worksheet starkstates.mw to see
how these states look like.
The only nontrivial secular equation for the L-shell problem then is (cf.
Eq. (2.32)):
−E (1) w12
det = 0
w12 −E (1)
In a similar fashion, one can study the Stark problem for the nine
M shell states. The perturbation matrix can be decomposed into five
blocks corresponding to the states with magnetic quantum numbers
m = −2 to m = 2.
Chapter 3
29
30
• EM potentials
the scalar potential Φ and the vector potential A are defined via
B = ∇×A (3.1)
∂A
E = −∇Φ − (3.2)
∂t
∂A
֒→ FL = q −∇Φ − + (v × (∇ × A)) (3.3)
∂t
this equation can be rewritten by introducing a generalized (i.e., velocity-
dependent) potential energy:
U := q(Φ − A · v) (3.4)
one can show that Eq. (3.3) can be written as
d
FL = −∇U + ∇v U (3.5)
dt
i.e.,
∂U d ∂U
FLi = − − i = 1, 2, 3
∂xi dt ∂ ẋi
the generalized potential paves the way to set up the Lagrangian:
• Lagrangian
m 2
L= T −U = v − qΦ + qA · v (3.6)
2
if one works out the Lagrangian equations of motion one finds ma = FL ,
i.e., Newton’s equation of motion with the Lorentzian force. This shows
that the construction is consistent. Now we are only one step away from
the Hamiltonian:
• Hamiltonian
H =p·v−L (3.7)
31
p = ∇v L = mv + qA (3.8)
1
v = (p − qA) (3.9)
m
if one uses Eq. (3.9) in Eq. (3.7) one arrives at
1
H= (p − qA)2 + qΦ (3.10)
2m
for details see [GPS], Chaps. 1.5 and 8.1
• add a scalar potential V (to account, e.g., for the Coulomb potential
of the atomic nucleus)
• q = −e
• quantization: p → p̂ = −i~∇
1
֒→ Ĥ = (p̂ + eA)2 − eΦ + V
2m
1
= (−~2 ∇2 − i~e∇ · A(r, t) − i~eA(r, t) · ∇ + e2 A2 (r, t))
2m
− eΦ(r, t) + V (r)
~2 2 e~ e~
= − ∇ + V (r) + A(r, t) · ∇ + (∇ · A(r, t))
2m mi 2mi
e2 2
+ A (r, t) − eΦ(r, t)
2m
≡ Ĥ0 + Ŵ (t) (3.11)
The only solution to Eq. (3.13) which is compatible with the requirement
that free EM waves are transverse is the trivial one Φ = 0. A monochromatic
(real) solution of Eq. (3.12) reads
with the unit vector π̂ being orthogonal to the wave vector k. This ensures
that indeed ∇ · A = 0. If one inserts (3.14) into (3.12) one obtains the
dispersion relation ω = ck. For details on the solutions of the free Maxwell
equations see [Jac], Chap 6.5.
Using the gauge conditions we arrive at the perturbation
e e2 2
Ŵ (t) = A(t) · p̂ + A (3.15)
m 2m
For weak fields one can neglect the A2 term. Ŵ = me A · p̂ is the usual
starting point for a perturbative treatment of atom-radiation interactions in
the semiclassical framework. The perturbation depends on time. We need a
time-dependent version of perturbation theory to deal with it.
assume that Ŵ (t ≤ t0 ) = 0
֒→ t ≤ t0 : Ĥ0 |ϕj i = ǫj |ϕj i j = 0, 1, . . .
assume |ψ(t0 )i = |ϕ0 i , (initial state) (3.18)
X i
ansatz : |ψ(t)i = cj (t)e− ~ ǫj t |ϕj i (3.19)
j
X
= cj (t)|ψj (t)i (3.20)
j
X i
֒→ i~ċk = λ e ~ (ǫk −ǫj )t cj hϕk |ŵ(t)|ϕj i (3.21)
j
֒→ transition probabilities ϕ0 −→ ϕk
note that X X
pk = hψ|ϕk ihϕk |ψi = hψ|ψi = 1
k k
as it should.
Ansatz for solution of Eq. (3.21): power series expansion
(0) (1) (2)
ck (t) = ck (t) + λck (t) + λ2 ck (t) + . . . (3.23)
insertion into Eq. (3.21) yields
(0) (1) 2 (2)
i~ ċk + λċk + λ ċk + . . .
X (0) (1) (2)
i
=λ cj + λcj + λ2 cj + . . . e ~ (ǫk −ǫj )t hϕk |ŵ(t)|ϕj i
j
34
֒→
(0)
λ0 : i~ċk = 0
(1)
X (0) i
λ1 : i~ċk = cj e ~ (ǫk −ǫj )t hϕk |ŵ(t)|ϕj i
j
(2)
X (1) i
2
λ : i~ċk = cj e ~ (ǫk −ǫj )t hϕk |ŵ(t)|ϕj i
j
Z t
(1) (1) i i ′
֒→ ck (t) − ck (t0 )
= − e ~ (ǫk −ǫ0 )t hϕk |ŵ(t′ )|ϕ0 i dt′ (3.25)
| {z } ~ t0
=0
(as λ = 0 at t = t0 )
accordingly:
Z Z ′
2 (2) 1 X t ′ t ′′ i (ǫk −ǫj )t′ i (ǫj −ǫ0 )t′′
λ : ck (t) =− 2 dt dt e ~ e~
~ j t0 t0
ǫk − ǫ0
with ωk0 = ωk − ω0 =
~
transition frequency
Wk0 = hϕk |Ŵ |ϕ0 i = λhϕk |ŵ|ϕ0 i
transition
Z t matrix element
1 ′
p0→k = 2| eiωk0 t Wk0 (t′ ) dt′ |2
~ t0
transition probability
the latter result can be worked out explicitly and is not surprising: it ex-
presses probability conservation.
36
Examples
(i) Slowly varying perturbation
Let’s assume that the perturbation is turned on very gently after t = t0 .
It might then stay constant for a while and/or is turned off equally
gently. This is to say that the time derivative of Ŵ is a very small
quantity for all times.
k 6= 0
Z
i t iωk0 t′
֒→ ck (t) = − e Wk0 (t′ ) dt′
~ t0
t Z t
i h 1 iωk0 t′ ′ 1 iωk0 t′ ′ ′
i
= − e Wk0 (t ) − e Ẇk0 (t ) dt
~ iωk0 t0 iωk0 t0
Ẇk0 ≈0 hϕk |Ŵ (t)|ϕ0 i iωk0 t
= − e
ǫk − ǫ0
i
X hϕk |Ŵ (t)|ϕ0 i i
= c0 (t)e− ~ ǫ0 t |ϕ0 i + e− ~ ǫ0 t |ϕk i
k6=0
ǫ0 − ǫk
Comments:
(i) The argument can be generalized to strong perturbations. The
general adiabatic approximation then results in the statement: if
the perturbation varies slowly with time, the system is found in
an eigenstate of Ĥ(t) at all times.
(ii) Adiabatic conditions are realized, e.g., if atom beams are directed
through slowly varying magnetic fields (→ Stern-Gerlach experi-
ment) and in slow atomic collisions. In the latter case the electrons
adapt to the slowly varying Coulomb potentials of the (classically)
moving nuclei and do not undergo transitions. They are in so-
called quasimolecular states during the collision and back in their
initial atomic states thereafter.
(iii) further reading: [Boh], Chap. 20; [Gri], Chap. 10
(ii) Sudden perturbation
V
0 t ≤ t0 ≡ 0
Ŵ (t) =
Ŵ t > t0
t
k 6= 0
Z
i t iωk0 t′
֒→ ck (t) = − e Wk0 (t′ ) dt′
~ 0
Z
hϕk |Ŵ |ϕ0 i t iωk0 t′ ′
= e dt
i~ 0
hϕk |Ŵ |ϕ0 i iωk0 t
= − e −1
~ωk0
֒→ transition probability
4|Wk0|2
p0→k (t) = |ck (t)|2 = f (t, ωk0) (3.30)
~2
sin2 ωk0
2
t
ωk0 →0 t
2
f (t, ωk0) = 2
−→ (3.31)
ωk0 4
38
t→∞ πt
f (t, ωk0) −→ δ(ωk − ω0 ) (3.32)
2
2πtt→∞
֒→ p0→k |Wk0|2 δ(ωk − ω0 )
−→
~
The sudden perturbation sounds academic, but it has an important
application (see later).
(note that Ŵ = Ŵ † )
Z t
1 ′
֒→ ck (t) = eiωk0 t Wk0 (t′ ) dt′
i~ 0
(
1 hϕk |B̂|ϕ0 i i(ωk0 +ω)t
= − e −1
~ ωk0 + ω
)
hϕk |B̂ † |ϕ0 i i(ωk0 −ω)t
+ e −1
ωk0 − ω
2π
if t ≫ ω
(i.e. ∆ω ≪ ω):
4|Bk0 |2 n o
p0→k (t) = f (t, ω k0 + ω) + f (t, ω k0 − ω) (3.34)
~2
2πt n o
t→∞ 2
−→ |Bk0| δ(ωk − ω0 + ω) + δ(ωk − ω0 − ω)
~
with Bk0 = hϕk |B̂|ϕ0 i and Eq. (3.31)
Comments:
e 0
stimulated emission (of energy)
h w
(only possible if ϕ0 is not the ground state)
e k
• ωk0 = +ω ⇐⇒ ǫk = ǫ0 + ~ω
e k
absorbtion (of energy)
h w
e 0
(ii) Since ϕ0 is not the ground state in the case of stimulated emission
it makes sense to change the notation and use ϕi for the initial
and ϕf for the final state. Transition frequencies are then denoted
as ωf i etc.
40
with
e ∗ −ik·r
B̂ = A e π̂ · p̂ (3.36)
2m 0
(v) Validity of first-order time-dependent perturbation theory
We have two conditions to fulfill to be able to apply our results
for the periodic perturbation:
• criterion to avoid overlap of the resonances
2π 2π
∆ω = ≪ω ⇔ t≫
t ω
• validity criterion on resonance
4|Bf i |2 |Bf i |2 2
pi→f = f (t, ω f i ± ω = 0) = t ≪1
~2 ~2
~
⇔ t ≪
|Bf i |
combine 2π 2π ~
−→ = ≪
ω |ωf i | |Bf i |
⇔ |∆ǫf i | ≫ |Bf i |
41
3.3 Photoionization
Let us elaborate on the case of absorption. If the energy, i.e., the field
frequency ω is high enough, the atom will be ionized. The final state of the
electron will then be a continuum state and this requires some additional
considerations, since such a state is not normalizable. This implies that
pi→f = |hϕf ||ψ(t)i|2 is a probability density rather than a probability. A
proper probability is obtained if one integrates over an interval of final states
a) Transitions into the continuum1 .
e k+ D e Z ǫf +∆ǫ
e k-D e
Pi→f := pi→f (ǫf ′ )ρ(ǫf ′ ) dǫf ′ (3.37)
ǫf −∆ǫ
e 0
with ρ(ǫf ′ ): density of states in interval [ǫf − ∆ǫ; ǫf + ∆ǫ]
Let’s be a bit more specific and calculate the density of states for free-particle
continuum states
1 i
ϕf (r) = hr|pi = exp( p · r)
[2π~]3/2 ~
dp
֒→ d3 p = p2 dpdΩp = p2 dǫf dΩp
dǫf
p
= 2m3 ǫf dǫf dΩp
Z Z p Z
2 3 3 2
֒→ 1 = |ψ(p)| d p = 2m ǫf |ψ(p)| dΩp dǫf
Z
= ρ(ǫf )pi→f (ǫf )dǫf
with p
ρ(ǫf ) = 2m3 ǫf
Let’s go back to the probability (3.37) and insert the first-order result
(3.34):
Z ǫf +∆ǫ n o
4
Pi→f = 2 |Bf ′ i |2 ρ(ǫf ′ ) f (t, ωf ′ i + ω) + f (t, ωf ′ i − ω) dǫf ′
~ ǫf −∆ǫ
Z ǫf +∆ǫ
abs 4 2
Pi→f ≈ 2 |Bf i | ρ(ǫf ) f (t, ωf ′ i − ω)dǫf ′
~ ǫf −∆ǫ
Z ∞
4 2 sin2 ( ω̃t
2
)
≈ |Bf i | ρ(ǫf ) dω̃
~ −∞ ω̃ 2
2π
= |Bf i |2 ρ(ǫf )t
~
R ∞ sin2 ( ω̃t )
where we have used −∞ ω̃2 2 dω̃ = πt 2
. A similar result is obtained for the
case of emission. One defines a transition rate wi→f = dtd Pi→f to obtain
2π
e,a
|Bf i |2 ρ(ǫf )
wi−→ f = (3.38)
~ ǫf =ǫi ±~ω
43
b) Dipole approximation
In a typical photoionization experiment the wavelength of the applied EM
field is large compared to the size of the atom. This allows for a simplification,
which is called the dipole approximation:
2π
≫ a0
if λ = ⇒ eik·r ≈ 1
k
(3.36) e ∗
֒→ Bfdip
i = A hϕf |π̂ · p̂|ϕi i
2m 0
this can be rewritten by using
im
p̂ = [Ĥ0 , r] (3.39)
~
p2
for Ĥ0 = 2m
+ V (r)
im
֒→ hϕf |π̂ · p̂|ϕi i = hϕf |Ĥ0 π̂ · r̂ − π̂ · r̂Ĥ0 |ϕi i
~
im
= (ǫf − ǫi )π̂ · hϕf |r|ϕi i
~
The matrix elements in questions are the well-known dipole matrix elements
(cf. Sec. 2). For π̂ = ẑ we have the standard selection rules ∆m = 0, ∆l =
±12 . They result in characteristic angular dependencies of photoionized elec-
trons, e.g., (see assignment # 3)
dip,ẑ πe2
wi−→ f = |A0 |2 (ǫf − ǫi )2 ρ(ǫf )|hϕf |z|ϕi i|2
2~3
∝ cos2 θ
expressions for stimulated emission and absorbtion, but it shows that there
is another process which cannot be described in our semiclassical framework:
spontaneous emission, i.e., the emission of a photon (and transition to a
lower-lying state) without any external EM field. So, let’s take a look at
field quantization.
Ĥ = Ĥ1 + Ĥ2 + Ŵ
Ĥ = ĤA + ĤF +Ŵ (3.40)
| {z }
=Ĥ0
ĤA is the Hamiltonian of the atom, ĤF the (yet unknown) Hamiltonian of
the field, and Ŵ the interaction. We aim at a description of the problem
within first-order perturbation theory. So we know which steps we have to
take:
Steps for a first-order TDPT treatment
• determine Ĥ0
p̂2 √
(a) ĤA = 2m
+V
(b) ĤF =?
45
√
ĤA |ϕj i = ǫj |ϕj i
ĤF |ξk i = εk |ξk i ? (3.41a)
֒→ Ĥ0 |Φjk i = ĤA + ĤF |ϕj i|ξk i
= (ǫj + εk ) |ϕj i|ξk i
= (ǫj + εk ) |Φjk i (3.41b)
While the first step (determining Ĥ0 ) is a physics problem, the second
one (solving Ĥ0 ’s eigenvalue problem) is a math problem.
• determine Ŵ
here it seems natural to start from the semiclassical expression (3.35)
and replace the classical vector potential by an operator that acts on
the degrees of freedom of the EM field
e
Ŵ = Â · p̂ (3.42)
m
Not surprisingly, it turns out that consistency with the form of ĤF
determines the form of Â
• apply FGR
the main issue here is to calculate the transition matrix elements
a) Construction of ĤF
Loosely speaking, Hamiltonians are the quantum analogs of classical energy
expressions. Let’s look at the energy of a classical EM field in a cube of
volume L3
Z
1
WEM = (E · D + B · H)d3 r
2 L3
Z
ǫ0 1
= (E2 + c2 B2 )d3 r c2 = (3.44)
2 L3 µ0 ǫ0
46
λ mode index
2π
kλ = L
(nλx , nλy , nλz ) nλi ∈ Z wave vectors and numbers
1 X 2 Pλ Pλ
֒→ WEM = ωλ (Qλ + i ) Qλ − i (3.51)
2 λ ωλ ωλ
1 X
= Pλ2 + ωλ2 Q2λ (3.52)
2 λ
Doesn’t this look familiar? The EM field has the algebraic form of a collection
of harmonic oscillators. We know how to quantize the harmonic oscillator,
so here is the quantization recipe:
1 X 2
֒→ WEM → ĤF = P̂λ + ω 2 Q̂2λ = ĤF† (3.56)
2 λ
The algebraic form of ĤF together with the commutation relations (3.55)
determine its spectrum:
X 1
En1 ,n2 ,... = ~ ωλ nλ + nλ = 0, 1, 2, ...
λ
2
would then correspond to the ground- and the excited-state levels of that
particle. However, it is not clear what kind of particle this should be, since
Qλ and Pλ do not correspond to usual position and momentum variables
(and actually there is no parabolic potential around).
The fact that the spectrum is equidistant allows for an alternate inter-
pretation, in which each mode is associated with nλ ’quanta’ 4 , each of which
carries the energy ~ωλ . In this interpretation a mode does not have a ground
state and excited states, but is more like a (structureless) container that can
accomodate (any number of) quanta of a given energy. The quanta are called
photons. At this point their only property is that they carry energy, but
we will see later that there is more in store. Note that the photon inter-
pretation is only possible because the spectrum of the harmonic oscillator is
equidistant!
nλ : occupation
P number of mode λ
N = λ nλ : total number of photons in the field
h i
†
⇒ b̂λ , b̂λ′ = δλλ′ (3.60)
h i
b̂λ , b̂λ′ = 0 (3.61)
h i
† †
b̂λ , b̂λ′ = 0 (3.62)
Usually, one uses a short-hand notation for the eigenstates and writes
|ψnλ i ≡ |nλ i. These states are called (photon) number or Fock states. Note
that |ψ0 i ≡ |0i is not a vector of zero length, but the (ground) state as-
sociated with the statement that there are no photons in the mode (but
energy E0 = ~ωλ /2). As eigenstates of a hermitian operator the |nλ i fulfill
an orthonormality relation
!
ĤFλ − 12
֒→ nˆλ |nλ i = nλ |nλ i n̂λ = (3.67)
~ ωλ
nˆλ |0i = 0|0i = 0 this is a real zero! (3.68)
The eigenvalue nλ is the number of photons in the mode. This justifies the
name occupation number operator for n̂λ .
Let’s operate with creation and annihilation operators on these photon
number states. To do this we need a few relations that can be proven without
difficulty.
h i
b̂λ , n̂λ′ = b̂λ δλλ′ (3.69)
h i
b̂†λ , n̂λ′ = −b̂†λ δλλ′ (3.70)
∢ n̂λ b̂λ |nλ i = b̂†λ n̂λ |nλ i + b̂†λ |nλ i
†
= (nλ + 1) b̂†λ |nλ i
51
֒→ combine:
√
⇔ α= nλ + 1
√
֒→ b̂†λ |nλ i = nλ + 1 |nλ + 1i (3.71)
b̂λ |0i = 0
Generalization:
The electronic matrix elements are the same as in the semiclassical frame-
work. We have discussed them in dipole approximation in Sec. 3.3. Let’s look
at the photonic matrix elements. Basically, they result in selection rules:
q
hn1 ... nλ ...| b̂λ |n1 ... nλ ...i = niλ hnf1 ... nfλ ...|ni1 ... niλ − 1 ...i (3.83)
f f i i
and similarly
q
hnf1 ... nfλ ...| b̂†λ |ni1 ... niλ ...i = niλ + 1 δnf ni δnf ni ...δnf ni +1 ... (3.85)
1 1 2 2 λ λ
1. the photon numbers in |Φi i and |Φf i differ exactly by one in one single
mode. Otherwise orthogonality will kill the transition amplitude. This
is to say that first-order perturbation theory accounts for one-photon
processes only.
2. dipole selection rules are fulfilled for the crucial mode (assuming that
eikλ ·r ≈ 1).
with the plus sign for absorption and the minus sign for emission.
e) Spontaneous emission
∢ |Φi i = |ϕi i|0, 0, . . .i
Even if there is no photon around in the initial state we can obtain a
nonvanishing transition matrix element by operating with a creation operator
on the vacuum state |0, 0, . . .i. This corresponds to the process of spontaneous
emission: an excited electronic state decays by (one-) photon emission even
though no external radiation field is present. One may say that it is the
zero-point energy in the given mode that triggers the transition.
For the investigation of the quantitative aspects of spontaneous emission
one uses FGR (3.38). The transition rate can be written as
s.e. 2π
wi−→f = |Wf i |2 ρ(ef ) (3.89)
~s
e2 ~3
Wf i = −i hϕf |e−ikλ ·r π̂λ · ∇|ϕi i (3.90)
2ωǫ0 m2 L3
ǫf − ǫi
with ω =
~
The density of states in FGR refers to the photon density in the final states.
To calculate it recall
2π λ λ λ
kλ = (n , n , n ) nλi ∈ Z
L x y z
3
∆N ∆nx ∆ny ∆nz L
֒→ ρ(k) = = =
∆V ∆kx ∆ky ∆kz 2π
55
• they can be created and annihilated (i.e. they don’t last forever)
using similar arguments as for the translation of WEM to Ĥf one obtains
X
P̂F = ~kλ n̂λ
λ
and X
P̂F |n1 , n2 , . . .i = ~kλ nλ |n1 , n2 , . . .i
λ
Brief introduction to
relativistic QM
Literature: [BD]; [BS], Chap. 1.b; [Mes], Chap. 20; [Sch], Chap. 13; [Lib],
Chap. 15; [SN], Chap. 8
The first reference is a classic textbook. The latter book chapters provide
condensed accounts on relativistic quantum mechanics.
E 2 = p2 c2 + m2 c4 (4.1)
• quantization : E −→ i~∂t
~
(standard rules) p −→ ∇
i
֒→ p2 −→ −~2 ∇2
֒→ E 2 −→ −~2 ∂t
1
In the following, m always denotes the rest mass m0 .
58
59
b) Discussion
6. Add Coulomb potential to free KGE and solve it (in spherical coordi-
nates)
−→ yields wrong fine structure of hydrogen spectrum (i.e., contradicts
experimental findings)
7. In 1934, the KGE was recognized as the correct wave equation for
spin-zero particles (mesons).
c~ X c~ X
֒→ −~2 ∂t2 Ψ = αj ∂xj + βmc2 αk ∂xk + βmc2 Ψ
i j
i k
n X ~ 3X o
2 2 2 2 4
= −~ c αj αk ∂xj ∂xk + mc αj β + βαj ∂xj + β m c Ψ
i j
jk
X αj αk + αk αj ~mc3 X
= −~2 c2 ∂x2j xk Ψ + β 2 m2 c4 Ψ + (αj β + βαj )∂xj Ψ
2 i j
jk
αj αk + αk αj = 2δjk (4.7)
αj β + βαj = 0 (4.8)
αj2 = β 2 = 1 (4.9)
0 1
with
cpz c(px + ipy ) c(px − ipy )
χ1 = 2
, χ2 = 2
, χ′1 = , χ′2 = −χ1
E + mc E + mc E + mc2
cpz c(px + ipy ) c(px − ipy )
ϕ1 = 2
, ϕ2 = 2
, ϕ′1 = 2
, ϕ′2 = −ϕ1
E − mc E − mc E − mc
note that all these ’small components’ approach zero for v ≪ c.
u(1) , u(3) are interpreted as ’spin up’
u(2) , u(4) are interpreted as ’spin down’ solutions
E e e
2
m 0 c
5 1 1 k e V fo r e
0
2
h o le
-m 0 c
D ira c s e a
Dirac’s interpretation(1930):
In the vacuum all negative energy states (in the Dirac sea) are occupied.
Hence, if electrons are present at E > mc2 they cannot ”fall down” into the
Dirac sea because of the Pauli principle (electrons are fermions).
On the other hand, one can imagine that it is possible to excite one
electron from the Dirac sea to E > mc2 . Such an excitation corresponds
to a hole in the Dirac sea, which can be interpreted as the presence of a
positively charged particle — an anti-particle (i.e. a positron). This process
— electron-positron pair creation — has indeed been observed, and also the
reversed process — destruction of electron-positron pairs and γ-ray emission
(the latter to balance the total energy).
In fact, the first experimental detection of positrons in 1932 was consid-
ered a strong proof of Dirac’s theory.
c) Throw in (classical) EM potentials
use the ’minimal coupling prescription’
~
p −→ ∇ + eA = p̂ + eA
i
E −→ i~∂t + eφ
(4.11)
n o
2
֒→ i~∂t Ψ = cα̂ · (p̂ + eA) − eφ + βmc Ψ (4.12)
Ze
φ=
4πǫ0 r
i
ansatz : Ψ(r, t) = Φ(r)e− ~ Et
(4.13)
yields
n Ze2 o
−→ cα̂ · p̂ + βmc2 − Φ(r) = EΦ(r) (4.14)
4πǫ0 r
this stationary DE can be solved analytically!
65
2
E = m 0 c
3 s3 p 3 d
n = 3
3 d 5 /2
2 s2 p 3 p 3 /2 3 d 3 /2
n = 2 3 s 1 /2 3 p 1 /2
2 p 3 /2
2 s 1 /2 2 p 1 /2
1 s 1 .8 x 1 0 -4
e V
n = 1
1 s 1 /2
S c h rö d in g e r D ira c
⇐⇒
c2
֒→ σ · p̂ σ · p̂ ϕ = (E − V (r) − mc2 )ϕ (4.23)
E − V (r) + mc2
• expand
c2 c2 1 1 ε − V (r)
2
= 2
= ε−V (r)
≈ 1−
E − V (r) + mc ε + 2mc − V (r) 2m(1 + 2mc2 ) 2m 2mc2
• use the following identity for Pauli matrices and arbitrary vector oper-
ators Â, B̂
(σ · Â)(σ · B̂) = Â · B̂ + iσ · (Â × B̂)
to obtain
(T1 + T2 + T3 )ϕ = (ε − V (r))ϕ
with
ε − V (r) p̂2
T1 = 1− (4.24)
2mc2 2m
~ 1 dV
T2 = (σ · l̂) (4.25)
4m2 c2 r dr
~ 1 1 dV
T3 = (r · p̂) (4.26)
i 4m2 c2 r dr
68
Interpretation of terms
• For the interpretation of T1 note that
p̂2
(ε − V (r))ϕ ≈ ϕ
2m
p̂2 p̂4
֒→ T1 = − ≡ T̂N R + ĤKE (4.27)
2m 8m3 c2
ĤKE represents the lowest-order relativistic correction to the kinetic
energy (as it appears — without hats — in a classical treatment).
• Introducing the spin operator ŝ = ~2 σ, which due to the properties
of the Pauli matrices fulfills the standard commutation relations of an
angular momentum operator4 , T2 is identified as the spin-orbit coupling
term
T2 ≡ ĤSO .
Hence, spin and spin-orbit coupling are automatically included in a
relativistic treatment (which is why some authors insist that electron
spin is a relativistic property).
• T3 is not hermitian. Consider its hermitian average
T3 + T3† 1 ~ 1 dV ~ 1 dV
T̄3 = = (r · p̂) − (p̂ · r)
2 8m2 c2 i r dr i r dr
2
~
= ∇2 V (r) ≡ ĤDarwin .
8m2 c2
The Darwin term doesn’t have a nonrelativistic or classical counterpart.
It is usually associated with the ”Zitterbewegung” (trembling motion)
of the electron due to the nonzero coupling of electrons and positions
(or: large and small components of the Dirac spinor) [BD].
Now apply perturbation theory to the problem
Ĥ = Ĥ0 + Ŵ (4.28)
2 2
p̂ Ze
Ĥ0 = T̂N R + V (r) = − (4.29)
2m 4πǫ0 r
Ŵ = ĤKE + ĤSO + ĤDarwin (4.30)
4
for a recap of angular momentum and spin operators consult a QM textbook, e.g.
[Gri], Chaps. 4.3, 4.4
69
mc2 1 3
∆E (1) = − (Zα) 4
1 − , (4.31)
2n3 j+ 2
4n
i.e., the same results as in Eq. (4.18), which shows the consistency of the
approach.
Chapter 5
Molecules
70
71
N
X e2
Vee = (5.7)
i<j
4πǫ0 |ri − rj |
Sub into Eq. (5.8), multiply everything by Ψ∗n (r, R), use Eq. (5.9) and inte-
grate over electronic coordinates choosing Ψ∗n (r, R) = Ψn (r, R) such that
to obtain
X
T̂nuc + Em (R) + ∆m (R) − E) φm (R) = Ĉmn (R)φn (R) (5.11)
n6=m
with definitions
D E
∆m (R) = Ψm (R) T̂nuc Ψm (R)
r
X ~2
Ĉmn (R) = α
(Aαmn (R) · ∇Rα + Bmn (R))
α
Mα
Hence, we arrive at two SEs; one for the electronic and one for the nuclear
motion:
T̂el + V (R) Ψm (r, R) = Em (R)Ψm (r, R) (5.13)
T̂nuc + Um (R) φm (R) = Eφm (R) (5.14)
Um (R) = Em (R) + ∆m (R) . (5.15)
The third equation provides the connection of the two SEs. This set of
equations is the standard framework for the discussion of molecular structure
and properties.
Discussion
"antibonding"
R
De E 0 (R)
"bonding"
• Re
Figure 5.1: Typical PESs for M = 2 (Em (R) = Em (R)) as functions of the
internuclear distance R.
(via the PESs). The fast electrons follow the slow nuclear motion adi-
abatically, i.e., without undergoing transitions.
A closer inspection of the BOA shows that the neglected terms ∆m (R)
and Ĉmn (R) are proportional to the smallness parameter 1/Mα .
2
!
P̂ R
Eφn (R) = + Un (R) φn (R)
2µ
2
!
P̂R2 Ĵ
= + + Un (R) φn (R) (5.16)
2µ 2µR2
M1 M2
with the reduced mass µ = M1 +M2
.
PnJ (R)
φn (R) = YJM (θ, ϕ)
R
֒→ radial equation:
2 2
~ d ~2 J (J + 1)
− + Un (R) + PnJ (R) = EPnJ (R) (5.21)
2µ dR2 2µR2
E
Re
R
De D0 D1
U0 (R)
Figure 5.2: Typical ground-state PES and harmonic oscillator model of vi-
brational motion. De = U0 (Re ) is the binding energy and D0 = De − ~ω2 e the
dissociation energy of the molecule.
Summary:
with
• E0 (Re ) is independent of µ
• Evib ∝ √1
µ
1
• Erot ∝ µ
require uniqueness:
∂
֒→ ˆlz φλ (ϕ) = −i φλ (ϕ) = λφλ (ϕ)
∂ϕ
h i
ˆ
⇒ Ĥ, lz = 0
78
However: h i
2
Ĥ, l̂ 6= 0
⇒ use λ to classify states
H atom l = 0 1 2 3
notation s p d f
+
H2 λ = 0 ±1 ±2 ±3
notation σ π δ ϕ
ψ(r) = ψ(ξ, η, ϕ)
= ψ(ξ, −η, ϕ + π)
= ψ(−r)
= ψg (r) gerade (even) parity
ψ(r) = ψ(ξ, η, ϕ)
= −ψ(ξ, −η, ϕ + π)
= −ψ(−r)
= ψu (r) ungerade (odd) parity
1
Discussion of exact PESs En (R) and ”correlation diagrams” Enel (R) = En (R) − R
Note that
2
Enel (R → 0) = − ”united-atom limit”
n2
1
Enel (R → ∞) = − 2 ”separated-atom limit”
2n
Since the Enel are continuous functions of R, each united-atom state (e.g.
1sσg ) is correlated with a separated-atom state (e.g. σg (1s)).
σg (1s)
σ u (1s)
σg (1s) + σ u (1s)
Figure 5.4: Sketch of the electron density distributions of the lowest states
of gerade and ungerade parity in the separated-atom limit and how they add
to form an atomic state, e.g., on the left centre
Enel (R → 0) AO MO AO Enel (R → ∞)
-2.0 1s 1σg 1s -0.5
-0.5 2p0 1σu 1s -0.5
-0.5 2p1 1πu 2p1 -0.125
-0.5 2s 2σg 2s -0.125
-0.222 3p0 2σu 2s -0.125
-0.222 3d0 3σg 2p0 -0.125
-0.222 3d1 1πg 2p1 -0.125
-0.222 3d2 1δg 3d2 -0.056
-0.222 3p1 3πu 3p1 -0.056
-0.222 3s 4σg 3s -0.056
-0.125 4f0 3σu 2p0 -0.125
Table 5.2: Energies (in atomic units) and united- and separated-atom limits
of a few MOs of H+2
81
82
83