Solutions
Solutions
Solutions manual
HORATIU NASTASE
Chapter 0
mv 2 3
Eph = + W = kB T + W ⇒
2 2
hν − W hν
T = ∼ , (0.3)
3kB /2 3kB
where in the last ∼ we used the assumption of W of the same order as hν. But
νvisible ∼ 4−8×1014 Hz , h ' 6.6×10−34 m2 Kg/s , kB ' 1.4×10−23 m2 Kg/(s2 K) ,
(0.4)
so
4 − 8 6.6
T ∼ × 1000K ∼ 7 − 14 × 1000K. (0.5)
3 1.4
3) The photon has momentum pγ = h/λ and energy Eγ = hc/λ, while the
2
electron has initially negligible
p momentum and (rest) energy me c , and after has
0 0 2 2 2
momentum pe and energy (pe c) + (me c ) .
Conservation of energy is then
2
2 0
p 0 2 hc hc 2
0 2 2 2
Eγ + me c = Eγ + (pe c) + (me c ) ⇒ (pe c) = − 0 + me c − (me c2 )2 ,
λ λ
(0.6)
and conservation of momentum is
p~0e = p~γ − p~0γ ⇒ p02 2 2 0 2 0 2
e c = (pγ c) + (pγ c) − 2pγ pγ c cos θ. (0.7)
Identifying the two ways of calculating p0e , we find
1 1 h 1 h
− 0 = 0
(1 − cos θ) ⇒ λ0 − λ = 2 sin2 θ/2. (0.8)
λ λ me c λλ me c
3
4 Chapter 0
However, this is not very useful, since the circular pendulum is a classical object,
so the angular momentum is very large, so l is well approximated by being contin-
uous. If however we have a quantum system of the same kind, like for instance a
precession motion in a quantum system, then it would be useful.
7) Light is a relativistic system, so we would need to use Quantum Field Theory.
Since we have Quantum Mechanics, in some sense yes, we should use a form of the
Schrödinger equation, but really only the Quantum Field Theory formalism does
it.
1 Chapter 1
˜
so that ~a0 · ~b = 0, then normalize it,
~˜
~b0 = qb . (1.4)
~˜b2
˜ ~b0
~c˜ = ~c˜ − (~b0~c˜) , (1.6)
~b02
˜ ˜
so that ~c˜ · ~b0 = 0, as well as ~c˜ · ~a0 = 0, and finally normalize it,
˜
~c˜
~c0 = q . (1.7)
˜
~c˜2
3) If
X X
 = |aiha0 | , B̂ = |bihb0 | , (1.8)
a,a0 b,b0
then
X X X
 · B̂ = |aiha0 |bihb0 | = ha0 |bi|ai hb0 |. (1.9)
a,a0 ,b,b0 b0 a,a0 ,b
5
6 Chapter 1
P
Moreover, in the matrix representation, inserting a complete set k |kihk| = 1,
X
(Â · B̂)ij = hi|Â · B̂|ji = ha0 |bihi|aihb0 |ji
a,a0 ,b,b0
X X
= hi|aiha |kihk|bihb0 |ji =
0
Aik Bkj . (1.10)
a,a0 ,b,b0 ,k k
With integrals (on wave functions), i.e. acting on |xi ⊗ |x0 i, we have
Z b Z b0
2
||F || = 2
|f (x)| dx |g(x0 )|2 dx0 . (2.5)
a a0
and the first case is the same as δx0 , the second is infinite.
4) The action of δ 00 (x − y) is found from the definition of the derivative of a
distribution:
∂ ∂ ∂ ∂ ∂2
h δ(x − y)|φi = −h δ(x − y)| φi = +hδ(x − y)| 2 φi. (2.8)
∂x ∂x ∂x ∂x ∂x
7
8 Chapter 2
Then we find Z
dxδ 00 (x − y)φ(x) = φ00 (y). (2.9)
where f (n) is the n’th derivative at zero, and is a real number for a real function.
Then
∞
h i† X f (n) n †
f (Â) = ( ) = f († ) = f (A) , (2.11)
n=0
n!
Ân
P∞
so is Hermitian. Examples are e = n=0 n! , sin Â, cos Â.
6) According to the previous exercise,
1) We need
(x−x1 )2 (x−x2 )2
− −
P1 = |ψ(x1 )|2 ∼ e P2 = |ψ(x2 )|2 ∼ e
2
σ1 2
σ2
, , (3.1)
and P = P1 + P2 . That is always true, provided that the particles are independent
(so P1 is independent of P2 ).
2) In H = αp2 + βq 2 + pf (q), the first two terms are already symmetrized
with respect to ordering, only the last term remains. If f (q) = n αn q n , the fully
P
3) If the spectrum extends a finite amount, then there is at least one accumu-
lation point, so En+1 − En → 0. If Ek+1 − Ek has a monotonic sign ∀k, then this
accumulation point is one of the boundaries of the spectrum, Emin or Emax . If
Emax = +∞ or Emin = −∞, that is not necessary, so there is no needed condition.
4) If there is a single spinless particle in a potential, the observables (commuting
operators) are energy (Ĥ) and positions xi (X̂i ), i = 1, 2, 3.
5) If Ĥ = x2 , the Schrödinger equation ~i ∂t ψ = Ĥψ , with
i
ψ(t, x) = e− ~ Et ψ̃(x) (3.5)
gives the time-independent Schrödinger equation Ĥ ψ̃(x) = E ψ̃(x), which doesn’t
have a solution.
6) If |xa,b i ≡ |2i ± |3i, then
|ai + |bi |ai − |bi
|2i = , |3i = . (3.6)
2 2
9
10 Chapter 3
Then
ρ̂Ĥ = λ1 |1ih1| + λ2 |2ih2| + λ3 |3ih3| = Ĥ ρ̂. (3.9)
Therefore [ρ̂, Ĥ] = 0, so ρ̂ doesn’t evolve in time.
4 Chapter 4
γi = σi ⊗ σa , i = 1, 2, 3 , γ4 = 1l ⊗ σb , (4.1)
where σa and σb are chosen among the σi , and differ from each other. Then, since
σi2 = 1l, we also have γi2 = σi2 ⊗ σa2 = 1l ⊗ 1l = 1l4 . Also, σa2 = σb2 = 1l means that
also γ42 = 1l. Moreover, since {σi , σj } = 2δij (so that also {σa , σb } = 0),
for ultrarelativistic particles we have t = L/c, and using units natural for the
corresponding quantities, we have
∆m2 (eV 2 )c4 L(km)
p2 (t) = sin2 (2θ) sin2 × 1.267 . (4.8)
E(GeV )
5) From (4.60) in text, we have
∆m2 E2 + E1
E2 − E1 = ~Ω ' , = Ē → E , (4.9)
2E 2
so the unitary operator becomes
∆m2 ∆m2 2
!
− iEt cos2 θei 4E~ t + sin2 θe−i 4E~ t 2i sin θ cos θ cos ∆m
4E~
t
U (t) = e ~
2 ∆m2 ∆m2
.
2i sin θ cos θ cos ∆m
4E~
t
sin2 θei 4E~ t + cos2 θe−i 4E~ t
(4.10)
6) The state is
|φi = C [|1i ⊗ |1i + |0i ⊗ |1i + a|1i ⊗ |0i + |0i ⊗ |0i] , (4.11)
so the normalization condition is
1
1 = hφ|φi = C 2 [1 + 1 + a2 + 1] ⇒ C = √ . (4.12)
3 + a2
A state is not entangled if we can write it as |φi ⊗ |ψi, for some |φi and |ψi. We
immediately note that for a = 1, we can write the state as
|φi = C [(|1i + |0i) ⊗ (|1i + |0i)] , (4.13)
so it is not entangled. In general, we can write the state as
|φi = C [(|1i
+ |0i) ⊗ |1i + (a|1i
+ |0i) ⊗ |0i]
a 1
= C (|1i + |0i) ⊗ |1i + |1i + |0i α|1i , (4.14)
α α
so we see that we need a/α = 1 and 1/α = 1 to have a non-entangled state, therefore
indeed a = 1 is the only such case.
7) Consider the total density matrix
ρ̂0 = |φihφ| , (4.15)
and its trace over the states in the B subsystem,
ρ̂ = TrB ρ̂0 = C 2 TrB [(|1i ⊗ |1i + |0i ⊗ |1i + a|1i ⊗ |0i + |0i ⊗ |0i) ·
· (h1| ⊗ h1| + h0| ⊗ h1| + ah1| ⊗ h0|)] . (4.16)
The trace over B means B h0|ρ̂0 |0iB + B h1|ρ̂0 |1iB , giving
ρ̂ = C 2 [(a|1i + |0i)(ah1| + h0|) + (|1i + |0i)(h1| + h0|)]
= C 2 (1 + a2 )|1ih1| + (1 + a)(|1ih0| + |0ih1|) + 2|0ih0| .
(4.17)
Then the probability to be in state 1 in A, independently of the state in B, is
1 + a2
P1 = h1|ρ̂|1i = C 2 (1 + a2 ) = . (4.18)
3 + a2
13 Chapter 4
which is just the Taylor expansion of a function around a point. But then H1
becomes
Z
H1 = dxψ ∗ (x)ψ(x + a) , (5.2)
we can rewrite H2 as
Z ∞ Z
d2
H2 = dα dxψ ∗ (x)e−α dx2 ψ(x) , (5.4)
0
Then the variations of the phase space objects under the canonical transformation
generated by F are
and momentum
∂L
p= = q̇ , (5.9)
∂ q̇
so with Hamiltonian
p2
H = pq̇ − L = + kq. (5.10)
2
The Poisson brackets of two functions, f and g, of phase space, are
∂f ∂g ∂f ∂g
{f, g}P.B. = − . (5.11)
∂q ∂p ∂p ∂q
1
Canonical quantization means (since {, }P.B. → i~ [, ]) the replacement
But
im 1 im i~t
− =− 2 − 1+ ⇒
2~t 2d 2~t md2
p t 2 p t 2
im x − m0 imx2 1 x − m0
p0 t
− i~t
+ = − i~t
− ip 0 x − , (6.9)
2~t 1 + md 2 2~t 2d2 1 + md 2 2m
so
p0 t 2
( )
1 x−
2 m p0 t
ψ(x, t) = q√ exp − 2 i~t
exp −ip0 x− . (6.10)
i~t
π d + md
2d 1 + md2
2m
Then
( 2 " #)
2 1 x − pm0 t 1 1
ρ = |ψ(x, t)| = q exp − i~t
+ i~t
i~t
π d + md
i~t
d − md
2d2 1 + md 2 1 − md2
" 2 #
1 x − pm0 t
=√ q exp − , (6.11)
~2 t2 d 2 1 + ~ 2 t2
π d2 + m 2 d2 m 2 d4
p2
4) Since H = 2m + αpx, we have that
[Ĥ, P̂ ] = αp̂ 6= 0 , (6.17)
so no, we cannot measure them simultaneously.
5) ∆E∆τ ≥ ~, but ∆E = 0.5eV , so
~
∆τ ≥ = τmin . (6.18)
0.5eV
6) The wave function is
(x−x0 )2 (x−x0 )2
− 2 − 2
2σ1 2σ2
C i p0 (x−x0 ) e e
ψ = C(ψ1 (x) + ψ2 (x)) = e~ √ + √ , (6.19)
π 1/4 σ1 σ2
R +∞
which we need to normalize, −∞
ψ ∗ (x)ψ(x) = 1, which leads to
1
2C 2 = −1/2 . (6.20)
1 1
σ12 + σ2
2 1 + √ 2
σ1 σ2
and
d2
Z
hp2 i = (−i~)2 dxψ ∗ (x) 2 ψ(x)
( dx
p0 + i~x p0 + i~x
Z +∞ 2 " #)
ψ22
∗ 2 2 ψ1 2 σ12 σ22
= dx ψ (x) p0 ψ(x) + ~ + 2 +~ x ψ1 + ψ2
−∞ σ12 σ2 σ12 σ22
Z +∞
2 2 2 ∗ ∗ ψ1 ψ2 2 ψ1 ψ2
= p0 + C ~ (ψ1 (x) + ψ2 (x)) 2 + 2 + i~x + 4 , (6.25)
−∞ σ1 σ2 σ14 σ2
~2
h(∆x)2 ih(∆p)2 i ∼ + ... , (6.27)
4
(6.28)
and finally
σ12 ~2 t2 σ2 ~2 t 2
2
h(∆x) i = C 2
1+ 2 2 + 2 1+ 2 2
2 m σ1 2 m σ2
21 Chapter 6
3/2
1 + 1
2 2 2 2
σ12 1+ ~ 2t 2 σ22 1+ ~ 2t 2
m σ1 m σ2
+ r . (6.29)
2
~ t 2 2
~ t 2
σ1 σ2 1 + m2 σ2 1 + m2 σ2
1 2
7 Chapter 7
~2 d2
− − V0 δ(x) ψ(x) = Eψ(x) , (7.1)
2m dx2
~2 0 2mV0
− [ψ (0+) − ψ 0 (0−)] − V0 ψ(0) = 0 ⇒ ψ 0 (0+) − ψ 0 (0−) = − 2 ψ(0). (7.2)
2m ~
√
For a bound state, E < 0, ~2 k 2 = (−E), so k = −E/~, and the wave functions
on both sides of the singularity must be decaying, so
and continuity of ψ(x) gives ψL (0) = ψR (0), so A = B. Then the condition (7.2)
implies
V0 m
k= . (7.4)
~2
√
The normalization condition, dx|ψ(x)|2 = 1 finally gives A = k. Then the
R
spectrum is
√
−E V0 m V 2 m2
k= = 2 ⇒E=− 02 . (7.5)
~ ~ ~
22
23 Chapter 7
Then
Z +L/2
2 nπx
m
hx in even = dx xm sin2
L L
−L/2
0 m odd
= 2 R +L/2
2πnx +L/2
L −L/2 dx x2k 1−cos2 L = 2 x2k+1 L2k+1
L 2(2k+1) − 2(πn)2k+1
I m = 2k ,
−L/2
(7.7)
and similarly
Z +L/2
m 2 nπx
hx in odd = dx xm cos2
L L
(−L/2
0 m odd
= L2k L2k
(7.10)
22k (2k+1)
+ (πn)2k+1
I m = 2k.
since one of the brackets is a sine, the other a cosine. On the other hand, if m = 2k,
Z +L/2 Z +L/2
m k 2 2k
hp in = dx|P̂ ψ| = ~ dx|ψ (k) (x)|2
−L/2 −L/2
nπ 2k Z +L/2 1 ± cos 2πnx
2k 2 L
=~ dx
L L −L/2 2
2k
nπ~
= = (2mEn )k , (7.14)
L
which is already its classical value.
5) For a cosine potential, V (x) = V0 cos(ax), we have −V0 ≤ V (x) ≤ V0 . For the
wave function, we have (at least) n − 1 nodes for the nth eigenfunction. That means
that if n is very large (as we will see we generically have), then the wave function
is highly oscillatory.
Even though @ limx→±∞ V (x), effectively we can define U± = V0 (the maximum
value of the oscillatory potential), since there will always be a place where V0 is
reached, no matter how far away in x we go. Moreover, for energies E < −V0 , there
is no solution, as usual, and for E > V0 , there is a continuum of solutions, also as
usual.
Naively, for −V0 < E < V0 , there will be a discrete spectrum, but! each local
minimum in the periodic potential will have its own discrete spectrum, and when
putting everything together, the various discrete spectra will interact, and the (ini-
tially identical) energy levels will split, obtaining energy bands.
6) To prove (7.83), we start from (7.82). Starting with the 4th equation, we
obtain
Q k2 − κ3
= e2ik2 b . (7.15)
P k2 + κ3
Then from the 3rd equation in (7.82), we obtain
16k1 k22 κ3
= 2 .
[(k2 + κ3 )(k1 + k2 ) + (k2 − κ3 )(k1 − k2 ) cos 2k2 (b − a)] + (k2 − κ3 )2 (k1 − k2 )2 sin2 2k2 (b − a)
(7.19)
To compare with the formula in the text, note that
ξKL = k2 (b − a)
4k1 k22 κ3
4ηζξ 2 =
K4
2 2
k (k1 + κ3 )
ξ 2 (η + ζ)2 = 2
K4
2
(k + k1 κ3 )2
(ξ 2 + ηζ)2 = 2 , (7.20)
K4
and the rest is algebra.
To check |R|2 + T = 1, from the value of R above, we find
|R|2
2
[(k2 + κ3 )(k1 − k2 ) + (k2 − κ3 )(k1 + k2 ) cos 2k2 (b − a)] + (k2 − κ3 )2 (k1 + k2 )2 sin2 2k2 (b − a)
= 2 ,
[(k2 + κ3 )(k1 + k2 ) + (k2 − κ3 )(k1 − k2 ) cos 2k2 (b − a)] + (k2 − κ3 )2 (k1 − k2 )2 sin2 2k2 (b − a)
(7.21)
and a bit of algebra proves that indeed |R|2 +√T = 1.
√ √
7) Defining, as usual, κ = U0 − or k = − U , and k1 = − U1 , the wave
functions for > U1 are:
√ √
ψ(x) = ei x + Re−i x , x < 0
Seik1 x , x > L
ψ(x) = (
Aeκx + Be−κx U1 ≤ ≤ U0
ψ(x) = ikx −ikx
(7.22)
Ce + De ≥ U0 .
a) U1 ≤ ≤ U0 . In this case, continuity of ψ(x) at x = 0 and x = L gives
1+R=A+B ⇒R=A+B−1,
κL −κL ik1 L −ik1 L+κL B −2κL
Ae + Be = Se ⇒ S = Ae 1+ e . (7.23)
A
Continuity of the log derivative of ψ at x = L gives
ik1
AeκL − Be−κL B 1−
κ κL −κL
⇒ = e2κL κ
ik1
, (7.24)
Ae + Be A 1+ κ
Then
2
B −2κL
T = |S|2 = e2κL |A|2 1 + e
A
√ 2
2
4e2κL κ
= √ 2 √ 2
k1 2 +k1 2
1− (1 − e−2κL ) +
κ κ (1 + e2κL )
1
= " √
2
#. (7.26)
√ 2 k
1− κ2 1
√ 2
+k
√ 1 2 +k1
2
+ sinh κL √
2 /κ
+ 2 √
and
(κ2 − )2 U02
= , (7.28)
4κ2 4(U0 − )
so we reproduce the result in the text.
b) > U0 . In this case, continuity at x = 0 and x = L gives
1+R=C +D ⇒R=C +D−1
ikL −ikL ik1 L i(k−k1 )L D −2ikL
Ce + De = Se ⇒ S = Ce 1+ e . (7.29)
C
Continuity of the log derivative of ψ at x = L gives
k1
CeikL − De−ikL D 1−
ik ikL −ikL
= ik1 → = e2ikL k
k1
, (7.30)
Ce − De C 1+ k
Then
2
D −2ikL
T = |S|2 = |C|2 1 + e
C
√ 2
4 2k
= √ 2 2 2 √ 2
k12
1 + kk1 + 1 − kk1 1 − k − 2 cos 2kL 1 −
1+ k k2 1− k2
27 Chapter 7
1
=
k2
. (7.32)
√ 2 1− k12 (1− k2 )
k1 +
√ 2
2
+ √ 2 sin kL
2 k
where
(k 2 − )2 U02
2
= , (7.34)
4k 4( − U0 )
reproducing the result in the text.
8 Chapter 8
At this moment, α is fixed in terms of β (and the given constants a, b, c), but β
is free. However, we still need to impose that
q12 + q22 = γ(q1 + αq2 )2 + δ(q1 − βq2 )2 , (8.6)
which fixes δ and γ, but also gives an extra constraint,
α 1 1
δ= γ,γ= α = ⇒
β 1+ β α2 1 + β
α
α
α2 + αβ = 1 + . (8.7)
β
2) For the Hamiltonian
N h i
â†n ân + αâ†n+1 ân + h.c.
X
H= , (8.8)
n=1
28
29 Chapter 8
so that finally
N
2πj
â†j âj 2 + 2α cos
X
H= . (8.12)
j=1
N
∗
hα1∗ |(â + ↠)3 |α2 i = [(α2 + α1∗ )3 + α2 + α1∗ ]hα1∗ |α2 i = [(α2 + α1∗ )3 + α2 + α1∗ ]eα1 α2 ,
(8.15)
where we have used the commutation relations and the fact that a|αi = α|αi and
hα∗ |a† = α∗ hα∗ |.
Then we divide the exponent we need to calculate in N equal parts, and insert a
completeness relation (as at exercise 3) after each such term, to obtain
N Z
dαi dαi∗
P
Y N ∗ i † † 3 i † † 3
I= e− i=1 αi αi |α1 ihα1 |e N [â â+β(â +â) ] |α2 ihα2∗ |e N [â â+β(â +â) ]
i=1
2πi
i † † 3
∗ ∗
...|αN ihαN |e N [â â+β(â +â)(] |αi
N N N
dαi dαi∗
Y Z X X
= |α1 i exp − αi∗ αi + αi∗ αi+1
i=1
2πi i=1 i=1
30 Chapter 8
N
)
1 X ∗ ∗ 3 ∗
+ [α αi+1 + β(αi + αi+1 ) + β(αi + αi+1 )] . (8.16)
N i=1 i
5) Defining as before
â + â†
r
mω
Q̂ = √ = x̂ , (8.17)
2 ~
we obtain
~ ~
hx̂2 in = hn|x̂2 |ni = hn|Q̂2 |ni = hn|(â2 + (↠)2 + â↠+ ↠â)|ni
mω 2mω
~ ~
= hn|(2↠â + 1)|ni = (2n + 1)
2mω
3/2 2mω
~
hx̂3 in = hn|Q̂3 |ni
mω
3/2
~
= hn|(â3 + (↠)3 + â(↠)2 + ↠â↠+ (↠)2 â + â2 ↠+ â↠â + ↠â2 )|ni
2mω
= 0. (8.18)
and, after redefining the sums in the various terms such that we have a common
factor y n , we get
∞
X p
y n [(n + 2)(n + 1)Cn+2 − 2 λ̃Cn−1 + 2˜
n Cn ] = 0. (8.29)
n=0
1) For a one-dimensional particle in potential V (x), from (9.13) in the text, we get
the evolution operator
i t 0 ~2 d 2
Z
Û (t, t0 ) = T exp − dt − + V (x) . (9.1)
~ t0 2m dx2
But for a Hamiltonian that is time-independent, like this one, we get
~2
h i
− ~i (t−t0 ) − 2m +V (x)
Û (t, t0 ) = e , (9.2)
i P̂ 2 P̂ 2 − i (t−t0 ) P̂ 2 P̂ 2
ĤH = Û † ĤS Û = e+ ~ (t−t0 ) 2m e ~ 2m = Ĥ
S = . (9.4)
2m 2m
Then the time evolution equation for a generic operator ÂH is (from eq. (9.28)
in the text)
dÂH ∂
i~ = [ÂH , ĤH ] + i~ ÂH . (9.5)
dt ∂t
We consider ÂH = X̂ and P̂ . In the first case, we obtain
dP̂H
i~ = 0 + 0 ⇒ P̂H (t) = P̂H (t0 ) = P̂S . (9.6)
dt
In the second case, we obtain (from (9.24) in the text)
In turn, that means that the evolution of the state in the Ŵ picture is
the interaction picture evolution equations for âI (t) and â†I (t) are
d
i~ âI (t) = [âI (t), Ĥ0 ] + 0 = Ŵ [â, ~ω(↠â + 1/2)]Ŵ −1
dt
= ~ωâI (t)
d †
i~ âI (t) = Ŵ [↠, ~ω(↠â + 1/2)]Ŵ −1 = −~ωâ†I (t). (9.15)
dt
The evolution of states is given by
i~∂t |ψI (t)i = Ĥ1,I |ψI (t)i = λŴ (â + ↠)3 Ŵ −1 |ψI (t)i. (9.16)
But
† †
Ŵ (↠+ â)3 Ŵ −1 = eiω(â â+1/2)(t−t0 ) (â + ↠)3 e−iω(â â+1/2)(t−t0 )
= (â + ↠)3 + [iω(↠â + 1/2), (â + ↠)3 ] + ...
= (â + ↠)3 − (a + a† ) + a3 − (a† )3 + a2 a† − (a† )2 a , (9.17)
so finally
i~∂t |ψI (t)i ' λ (â + ↠)3 − (a + a† ) + a3 − (a† )3 + a2 a† − (a† )2 a |ψI (t)i. (9.18)
1) eiS = cos S + i sin S is highly oscillatory, so the (path) integral over it is not
well defined. To make sense, we must replace it with an honest Gaussian, e−SE , by
a ”Wick rotation” to Euclidean space. If iS = −SE , we see that we can achieve it
R t = −itE , Rsince in the Lagrangian L = T − V , the potential term becomes
with
i dt(−V ) = − dtE V , and if V ≥ 0 (as needed in well defined quantum systems),
then we have a decreasing exponential, as we wanted.
2) For a term Ĥ1 = α(P̂ X̂ 2 + X̂ P̂ X̂ + X̂ 2 P̂ ), in order to be able to use the same
formula (10.20) in the text,
That means that we need to replace Ĥ1 → Ĥ1 + 3αi~X̂ everywhere in the fol-
lowing.
3) For the Hamiltonian H(p, x) = 34 p4/3 + V (x), the path integral in phase space
formula, (10.21), remains valid,
( Z 0 )
Z Z t
0
U (t , t0 ) = Dx(t) Dp(t) exp i dt[p(t)ẋ(t) − H(p(t), x(t))] . (10.3)
t0
The first observation is that the equation of motion for p in the exponent is
1/3
p (t) = ẋ(t), which when substituted back in the exponent gives the classical
Lagrangian,
1
L = ẋ4 − V (x). (10.4)
4
That already tells us that the first approximation to the path integral is still the
classical action, eiScl [x,ẋ] . Next, we see that indeed
3
− p4/3 + pẋ − V
4
3 3 ẋ4
= − (p1/3 − ẋ)4 − 2ẋ(p1/3 − ẋ)3 − ẋ2 (p1/3 − ẋ)2 + −V , (10.5)
4 2 4
so that we can write
Z R h 4 iZ
0 i dt ẋ4 −V
U (t , t0 ) = Dx(t)e Dp(t)
Z
3 1/3 4 1/3 3 3 1/3 2
exp i dt (p − ẋ) − 2ẋ(p − ẋ) − (p − ẋ) (. 10.6)
4 2
35
36 Chapter 10
It would seem that this is fine, since the first exponent is eiS , but the path
integral over p(t), even shifted by ẋ, still has coefficients that depend on ẋ, so gives
an (uncalculable) function of ẋ, which means that the integrations over fluctuations
around the classical solution give a result that differs from the one in the standard
eiS path integral, and we cannot calculate them.
4) For the Hamiltonian
when going to the harmonic phase space path integral, we still have eq. (10.41),
∗
hα∗ |Ĥ(a† , â)|pi = H(α∗ , p)eα β
, (10.8)
U (α∗ , t0 ; α, t) ( Z 0 )
i t
Z Z ∗
∗ α̇ (τ ) ∗ ∗
= Dα(t) Dα (t) exp dτ α(τ ) − H(α (τ ), α(τ )) + α (t)α(t) .
~ t i
(10.9)
For the 3-point function, since there are only pairs of J’s, when putting J = 0,
we obtain 0, so
G3 (t1 , t2 , t3 ) = 0. (10.12)
Defining
d2
∆−1 (t1 , t2 ) = i + ω 2
δ(t1 − t2 ) (10.15)
dt21
dtJ(t)q(t) ≡ J · q, dt1 dt2 q(t1 )∆−1 (t1 , t2 )q(t2 ) = q · ∆ · q, we write
R R R
and
Z
1 −1
Z[J] = Dq exp − q · ∆ · q + iJ · q
Z 2
1 −1 1
= Dq exp − (q − i∆ · J) · ∆ · (q − i∆ · J) − J · ∆ · J
Z2 Z 2
1
= N exp − dt1 dt2 J(t1 )∆(t1 , t2 )J(t2 ) , (10.16)
2
which is the result we wanted, and N contains the result of the Gaussian path
integration over q(t) − i dt0 ∆(t, t0 )J(t0 ).
R
2
To find ∆(t1 , t2 ), we Fourier transform ∆−1 (t1 , t2 ) ≡ i dt
d
2 + ω 2
δ(t1 − t2 ), to
1
obtain
0
dp e−ip(t−t )
Z
∆(t1 , t2 ) = i . (10.17)
2π p2 − ω 2
7) For the partition function
Z Z
1
Z[J(t)] = N exp − dt dt0 J(t)∆(t, t0 )J(t0 )
Z Z2 Z Z
+λ dt1 dt2 dt3 dt4 J(t1 )J(t2 )J(t3 )J(t4 ) , (10.18)
−2L/r2
Z
dr q + θ − θ0 = 0 ⇒ θ = θ(r). (11.19)
L2
2m E + Br −
r2
Since we are not interested in the trajectory r(t), but only in the deflection angle,
coming from θ = θ(r), we ignore the first equation, and concentrate on the second,
which is, explicitly,
Z r
2L
θ(r) = dr0 q . (11.20)
2
r02 2m E + rB0 − rL02
The particle comes from r = ∞, down to the inflection point, the minimum of
θ(r), so where the denominator in θ(r) vanishes, called θ1 (r0 ), then back to r = ∞.
We then obtain
s 2
L2 B B
r0 = + − . (11.21)
2m 2 2
In conclusion, we have
Z ∞
2L
θ1 = dr0 q , (11.22)
B L2
r0 r02 2m E + r0 − r 02
and, drawing perpendiculars onto the angle lines corresponding to the two limits,
we obtain that θ1 − π/2 is half the deflection angle δα, so
δα = 2θ1 − π. (11.23)
We can calculate the integral in θ1 , defining first 1/r = x, then x − mB/L2 = y,
to obtain
1
r0 − mB
L2
Ly
θ1 = 2 arcsin q
m2 B 2
2mE + L2 − mB
L2
1 mB mB
r0 − L2 L 2
= 2 arcsin q + 2 arcsin q . (11.24)
m2 B 2 m2 B 2
2mE + L2 2mE + L2
i
4) For the wave function ψ = Ae ~ S(~r,t) , the modified Hamilton-Jacobi equation
in eq. (11.27) in the text is
~ 2
(∇S) ~ 2A
~2 ∇
+ V (~r) + ∂t S − =0, (11.25)
2m 2m A
and when substituting S(~r, t) = W (~r) − Et, to obtain eq. (11.41) in the text, and
adding the continuity equation, we obtain the coupled system of equations for W (~r),
A(~r, t),
1 ~ ~ 2A
~2 ∇
(∇W )2 + V (~r) − E − =0
2m 2m A
1 ~ ~ 1 ~2
∂t A = − ∇A · ∇W + A∇ W . (11.26)
m 2
41 Chapter 11
This system can’t be uncoupled in general, so it stands for the equation for A.
~2 ∇~ 2A
However, in the geometric optics approximation, 2m A → 0, so we can solve
the first equation for W and replace it in the second, which now becomes a single
equation for A.
If moreover V (~r) = 0 (for a free particle), we obtain
√
W = 2m(~ p · ~r + const.) , (11.27)
so the equation for A is
1 ~2
p
∂t A = − ~ ⇒ A = Cei p~·~x− 2m t .
p~ · ∇A (11.28)
m
5) For a one-dimensional harmonic oscillator, V (x) = kx2 /2, the WKB approxi-
mation can be valid, yes, if the usual geometrical optics approximation is valid,
δλ λ
1, (11.29)
λ
which means: far from the turning points with E = V (r); also it gets better for
larger n (states are more classical), in which case E − En becomes also smaller.
6) For the potential V (x) = −B/x, the WKB wave function is
" s #
i x 0
Z
const. B iEt
ψ(x, t) = 1/4 exp ± ~ dx 2m E +
x
− . (11.30)
E+B x
x0 ~
The domain of validity is, as always, the domain of validity of the geometric
optics approximation,
δλ λ 1 δλ (E − V (x))
1⇒ 1. (11.31)
λ 2 E − V (x)
p
Then for E > 0, at large x, λ = ~/ 2m(E + B/x) is approximately constant,
so this is satisfied. For E < 0, we need to be far from the turning points, where
E + B/x = 0.
7) In one dimension, the equations for W and A (as written at exercise 4) become
1 ~2 A00
W 02 = E − V (x) +
2m 2m A
1 2 0 0
Ȧ = − (A W ) . (11.32)
2mA
In order to find the first order corrections to the WKB approximation, we can put
the 0th order solution (the WKB one) on the right-hand side of the first equation
above, and the resulting solution is also used in the second equation, to find the
next order of the coefficient A. Therefore, from the WKB solution,
const. A00 5 V 02 1 V 00
A= ⇒ = − , (11.33)
[E − V (x)]1/4 A 16 (E − V (x))2 4 E − V (x)
and this A00 /A is used in the first equation, obtaining
r
~2 A00
Z
W = dx E − V (x) +
2m A
42 Chapter 11
s
~2 V 02 V 00
Z
5 1
= dx E − V (x) + − . (11.34)
2m 16 (E − V (x))2 4 E − V (x)
Next, we use the 0th order (WKB) A and the 1st order W to obtain A2 W 0 to on
the right-hand side of (11.32), so
v
V 02 V 00
u 5 1
2 2 − 4 E−V (x)
u
~ 16 (E−V (x))
A2 W 0 = const. 1 +
t
⇒
2m E − V (x)
u v 0
5 V 02 1 V 00
const.
Z u
~ 2 16 (E−V (x)) 2 − 4 E−V (x)
dt[E − V (x)]1/4 1 +
t
A=− .
2m 2m E − V (x)
(11.35)
12 Chapter 12
˙ 2 is invariant.
such that q 2 + q̃ 2 is invariant, and |q̇|2 + |q̃|
q 2πik
-If is multiplied by a Z4 element, e 4 , for k = 0, 1, 2, 3 (so both q and q̃
q̃
are multiplied by the same element), then (q 2 + q̃ 2 )2 is invariant, and so are |q̇|2
˙ 2.
and |q̃|
-Also an invariance is the Z2 that exchanges q and q̃, and the Z2 that acts only on
q as q → −q. All other symmetries are composites of these (or included in these).
For the continuous SO(2) symmetry, the representation is the fundamental, or
defining, one. For the Z4 and the Z2 ’s, the representation is the fundamental, or
defining, one.
2) In general, the Noether charge is
∂L j
Qa = (iTa )i qj . (12.3)
∂ q̇i
For the case of the SO(2) symmetry at exercise 1, qi refers to q, q ∗ , q̃, q̃ ∗ . The
complex conjugates can be dealt with by adding the complex conjugate of the result,
so we obtain
m ∗ ˙∗ q
Qa = q̇ q̃ (iTa ) + h.c. (12.4)
2 q̃
But for SO(2), for an infinitesimal angle θ = , so that we obtain the generator
(iTa ), we find
cos θ sin θ 0 1 0 1
= 1l + + O(2 ) ⇒ iTa = , (12.5)
− sin θ cos θ −1 0 −1 0
SO(2)) symmetry q → eiα q. There are no particular discrete symmetries, since the
symmetries of q 3 differ than the ones of q 5 .
-in the case I, for V = (1/q 3 + 1/q 5 ) + (1/q̃ 3 + 1/q̃ 5 ), for the same reason, there is
no discrete symmetry for each of the two brackets. There is only the Z2 symmetry
exchanging q and q̃. This potential also doesn’t admit any continuous symmetries
(the SO(2) symmetry was for a potential of the type V = V (q 2 + q̃ 2 ) only ).
6) a) We identify the two matrices with the generic matrix elements as
1 0 0 0 0. 1
A = 0 1 0 = e B = 0 1 0 = a , (12.18)
0 0 1 1 0 0
then indeed they satisfy the multiplication table of Z2 , where e corresponds to +1
and a to −1, so:
e · e = e , e · a = a , a · a = e. (12.19)
We can indeed check that A2 = A, A · B = A, B 2 = A.
b) The representation for A and B is not block-diagonal, so it is irreducible.
c) For a regular representation, D(g1 )|g2 i = |g1 g2 i. Since we have D(e) = A and
D(a) = B, we would need
D(e)|ei = |e · ei = |ei , D(e)|ai = |e · ai = |ai ,
D(a)|ei = |a · ei = |ai , D(a)|ai = |a · ai = |ei. (12.20)
But we must then identify |ei as the first basis element, and |ai as the second basis
element of the column vector. We see then that for A and B as above, it doesn’t
1 0
work. In fact, from the above conditions we see that we must have A =
0 1
0 1
and B = , for a regular representation.
1 0
d) Since we have a different dimensional space, the representation is not equiva-
lent to the roots of unity representation (±1).
7) In order to still have D(g1 )|g2 i = |g1 g2 i, we see that we need to have N × N
matrices. By comparison with the roots of unity representation, we see that if
2πi
g1 = e N , the matrix D(g1 ) takes |ei → |e2 i, |e2 i → |e3 i,... (shift by one), D(g2 )
shifts by two, etc. So:
0 0 ... 0 1 0 1 0 ... 0
1 0 ... 0 0 0 0 1 ... 0
D(e) = 1lN ×N , D(e2 ) = .. .. ... .. .. , ..., D(eN ) =
.. .. .. ... .. .
.. .. ... .. .. 0 0 0 ... 1
0 0 ... 1 0 1 0 0 ... 0
(12.21)
13 Chapter 13
1) For the harmonic oscillator, H(−x) = H(x), since we have only quadratic terms,
and in quantum mechanics [π̂, Ĥ] = 0. That means that both the three dimensional
and the one dimensional harmonic oscillator are parity invariant. The parity of a
state n of the one dimensional harmonic oscillator is
so is clearly parity invariant, being quadratic (it is even positive). The Hamiltonian
is time reversal invariant, having only quadratic terms, which means ρn is real (it
has no phase), and it is time-reversal invariant and parity invariant.
3) For the one-dimensional Hamiltonian
p21 p2
H= + 2 + V (|q1 |) + V (|q2 |) + V12 (|q1 − q2 |) , (13.4)
2 2
possible continuous internal symmetries would act only on (q1 , q2 ), but there are
none, since for SO(2), we would need q12 + q22 .
4) For the algebra [A, B] = C, [B, C] = A, [C, A] = B, we need to check the
Jacobi identities to see if it is a Lie algebra. There is only one identity,
[C, [A, B]] + [A, [B, C]] + [B, [C, A]] = 0 , (13.5)
so we obtain (since 1 = A, 2 = B, 3 = C)
0 0 0 0 0 −1 0 1 0
TA = −i 0 0 1 , TB = −i 0 0 0 , TC = −i −1 0 0 .
0 −1 0 1 0 0 0 0 0
(13.8)
46
47 Chapter 13
~2 + S
5) For H = α(S ~ 2) + βS
~1 · S
~2 , we have
1 2
~ −1 = −S
T ST ~ ⇒ TS
~ 2 T −1 = S
~2 , T S
~1 · S
~2 = S
~1 · S
~2 , (13.9)
so the Hamiltonian is time-reversal invariant. Under parity spin is invariant, so the
Hamiltonian is explicitly invariant.
6) SU (N ) has N 2 − 1 independent components, meaning that there are N 2 −
1 generators Ta , so the dimension of the adjoint representation is N 2 − 1. For
equivalence, we would need to same dimension for the fundamental representation,
of dimension N , so N 2 − 1 = N , which has no solutions.
7) For the Lagrangian
N N
!
X |q̇i |2 X
2
L= −V |qi | , (13.10)
i=1
2 i=1
p
1) Considering that αjm = (j + m)(j − m + 1), for spin one matrices we obtain
(from eq. (14.76) in the text), for m = −1, 0, +1,
√
0 1 0
~ ~ 2
(J1 )mm0 = (α1,m+1 δm0 ,m+1 + α1,m δm0 ,m−1 ) = 1 0 1
2 2
0 1 0
√
0 1 0
~ ~ 2
(J2 )mm0 = (α1,m+1 δm0 ,m+1 − α1,m δm0 ,m−1 ) = −1 0 1
2i 2i
0 −1 0
−1 0 0
(J3 )mm0 = m~δmm0 = ~ 0 0 0 . (14.1)
0 0 +1
In the adjoint representation, given that the Lie algebra is [Ji , Jj ] = iijk Jk , we
have
c
(Ta )b = ifab c = −abc , (14.2)
We obtain 4 equations
θ φ+ψ
cos cos = cos |~
α/2|
2 2
θ φ+ψ α3
cos sin = sin |~
α/2|
2 2 |~
α|
θ φ−ψ α1
sin cos = sin |~
α/2|
2 2 |~
α|
θ φ−ψ α2
sin sin = sin |~
α/2| , (14.6)
2 2 |~
α|
leading to the equalities
α2 φ−ψ
= tan
α1 2
α3 φ+ψ
tan |~
α/2| = tan
|~
α| 2
2
α θ
α/2| 1 − 32 = sin2 ,
sin2 |~ (14.7)
|~
α| 2
and the first two equations can be solved for the Euler angles φ and ψ as
+α
α1 +
2 α3
α| tan |~
|~ α/2| −α
α1 +
2 α3
α| tan |~
|~ α/2|
tan φ = α2 α3 , tan ψ = α2 α3 . (14.8)
1− α| tan |~
α1 |~ α/2| 1+ α| tan |~
α1 |~ α/2|
3) Identifying the general matrix with the form with Euler angles, we obtain
φ+ψ φ−ψ
!
cos θ/2ei 2 i sin θ2 ei 2
a b
A= = φ−ψ φ+ψ , (14.9)
c d i sin θ2 e− 2 cos θ/2e− 2
which results in
θ
ad = cos2 θ
2 bc = − sin2
2
ab ac
a
d = ei(φ+ψ) , b
c = ei(φ−ψ) ⇒ e2iφ = , e2iψ = . (14.10)
cd bd
4) For the Lagrangian
where in the last equation we have used the fact that Tr σi = 0. Moreover, under the
SU (2) transformation q → U q, the kinetic term q̇i† q̇i is invariant as usual, whereas
the potential term is also invariant,
X X X
q † Aq = ai q † σi q → ai q † U † σi U q = ai q † σi q. (14.13)
i i i
50 Chapter 14
α · ~σ )2 = α
1) Since σ(i σj) = δij , so (~ ~ 2 , we obtain
|~
α| |~
α| |~
α| |~
α|
αi σi e 2 + e− 2 ~ · ~σ e
α 2 − e− 2 |~
α| α · ~σ |~
α|
e 2 = + = cosh + sinh , (15.1)
2 |~
α| 2 2 |~
α| 2
so that
αi σi |~
α|
Tr e 2 = 2 cosh . (15.2)
2
2) The decomposition is 1 ⊗ 1 = 0 ⊕ 1 ⊕ 2 (with dimensions 3 · 3 = 1 + 3 + 5), so
the angular momentum addition formulas in terms of Clebsch-Gordan coefficients
are
X
|11; 0mi = |11; m1 m2 ih11; m1 m2 |11; 0mi
m1 ,m2
X
|11; 1mi = |11; m1 m2 ih11; m1 m2 |11; 1mi
m1 ,m2
X
|11; 2mi = |11; m1 m2 ih11; m1 m2 |11; 2mi. (15.3)
m1 ,m2
the phase convention is that all Clebsch-Gordan coefficients are real, while
h1, 1; 1, m2 |1, 1; 1, 1i is real and positive, and the normalization condition is
X
(h1, 1; m1 , m2 |1, 1; jmi)2 = 1. (15.5)
m1 ,m2
51
52 Chapter 15
Considering (for instance) m01 = m02 = 0 and both signs in the recursion condi-
tion, we obtain
h1, 1; 0, ∓1|1, 1; 0, 0i = −h1, 1; ∓1, 0|1, 1; 0, 0i. (15.8)
Considering m02 = 1, m01 = −1 and the lower sign, we obtain
h1, 1; −1, +1|1, 1; 0, 0i = −h1, 1; 0, 0|1, 1; 0, 1i , (15.9)
and considering m01 = −, m02 = −1 and the lower sign, we obtain
h1, 1; +1, −1|1, 1; 0, 0i = −h1, 1; 0, 0|1, 1; 0, 10i. (15.10)
Then considering m02 = 1, m01 = 0 and the lower sign, we obtain
h1, 1; 1, 1|1, 1; 0, 0i = 0 , (15.11)
considering m02 = 0, m01 = −1 and the lower sign, we obtain
h1, 1; −1, −1|1, 1; 0, 0i = 0 , (15.12)
considering m01 = −1, m02 = 1 and the lower sign, we obtain
h1, 1; 0, 1|1, 1; 0, 0i = 0 , (15.13)
and considering m01 = 1, m02 = −1 and the upper sign, we obtain
h1, 1; 0, −1|1, 1; 0, 0i = 0. (15.14)
That means that we have only 3 nonzero Clebsch-Gordan coefficients, and the
normalization condition becomes
3(h1, 1; 0, 0|1, 1; 0, 0i)2 = 1. (15.15)
Given that the phase convention is that h1, 1; +1, −1|1, 1; 0, 0i is real and positive,
we finally obtain that
1
√ = h1, 1; +1, −1|1, 1; 0, 0i = h1, 1; −1, +1|1, 1; 0, 0i = −h1, 1; 0, 0|1, 1; 0, 0i
3
(15.16)
and the rest are zero.
4) For the addition 1 ⊗ 1 ⊗ 1 = 2 (total angular momentum 2), we have (in the
3 ways of decomposing the double sum)
|1, 1,X
j12 , 1, 2, mi
= |1, 1, 1; m1 , m2 , m3 ih1, 1, m1 , m2 |j12 , m12 ihj12 , 1; m12 , m3 |2, mi
m1 ,m2 ,m12 ,m3
|1, 1,X
1, j23 , 2, mi
= |1, 1, 1; m1 , m2 , m3 ih1, 1, m2 , m3 |j23 , m23 ih1, j23 ; m1 , m23 |2, mi
m1 ,m2 ,m23 ,m3
|1, 1,X
j13 , 1, 2, mi
= |1, 1, 1; m1 , m2 , m3 ih1, 1, m1 , m3 |j13 , m13 ih1j13 ; m2 , m13 |2mi ,
m1 ,m2 ,m3 ,m13
(15.17)
53 Chapter 15
where necessarily a†j,+ is either a†1,+ or a†2,+ (identifying the up spin 1/2 creation
operators on both sides), and a†j,− is either a†1,− or a†2,− (idem for the down spin
1/2).
Therefore the Clebsch-Gordan coefficients are
h1, 1; m1 , m2 |1, 1; j, mi
† †
(a1,+ )1+m1 (a1,− )1−m1 (a2,+ )1+m2 (a2,− )1−m2 (aj,+ )j+m (aj,− )j−m
= h0| p p p |0i.
(1 + m1 )!(1 − m1 )! (1 + m2 )!(1 − m2 )! (j + m)!(j − m)!
(15.21)
6) For the Lie algebra
[Ji , Jj ] = iijk Jk , [Ki , Kj ] = iijk Jk , [Ji , Kj ] = iijk Kk , (15.22)
we can make the redefinition
Ji + Ki Ji − Ki
Qi = , Pi = , (15.23)
2 2
after which we obtain
[Qi , Qj ] = iijk Qk , [Pi , Pj ] = iijk Pj , [Qi , Pj ] = 0 , (15.24)
so two commuting copies of SU (2). That means we can define two copies of the
Schwinger oscillators, representing the algebra as
Q− = a†1,+ a1,− , Q+ = a†1,− a1,+ , 2Qz = a†1,+ a1,+ − a†1,− a1,−
P− = a†2,+ a2,− , P+ = a†2,− a2,+ , 2Pz = a†2,+ a2,+ − a†2,− a2,− . (15.25)
The representations are defined by the double (j1 , j2 ), for representations of the
two independent SU (2)’s, and are written in terms of the Schwinger oscillators as
(a†1,+ )j1 +m1 (a†1,− )j1 −m1 (a†2,+ )j2 +m2 (a†2,− )j2 −m2
p p |0i. (15.26)
(j1 + m1 )!(j1 − m1 )! (j2 + m2 )!(j2 − m2 )!
7) For a composition of 4 angular momenta of j = 1, we have
1 ⊗ 1 ⊗ 1 ⊗ 1 = (0 ⊕ 1 ⊕ 2) ⊗ (0 ⊕ 1 ⊕ 2)
54 Chapter 15
=0⊗0⊕0⊗1⊕0⊗2⊕1⊗0⊕1⊗1⊕1⊗2⊕2⊗0⊕2⊗1⊕2⊗2
= 3 · 0 ⊕ 6 · 1 ⊕ 6 · 2 ⊕ 3 · 3 ⊕ 4. (15.27)
Indeed, we can check that the dimensions work out, since (the dimension of the
spin j representation is 2j + 1)
81 = 34 = 3 + 6 · 3 + 6 · 5 + 3 · 7 + 9. (15.28)
16 Chapter 16
(L21 + L22 + L23 )(eimφ ylm (θ)) = ~2 l(l + 1)eimφ ylm (θ) , (16.7)
55
56 Chapter 16
so that
l
X l
X X
0 ∗
Ylm (~u )Ylm (~v 0 ) = ∗
Um0 m Um 00 m Ylm0 (~
∗
u)Ylm 00 (~
v ). (17.9)
m=−l m=−l m0 ,m00
57
58 Chapter 17
Pl ∗
But U is unitary, so m=−l Um0 m Um 00 m = δm0 m00 , meaning
l
X l
X
Ylm (~u0 )Ylm
∗
(~v 0 ) = ∗
Ylm0 (~u)Ylm 0 (~
v) , (17.10)
m=−l m0 =−l
~ ×B
~τ = µ ~ , (17.13)
~ = 21 dV(~r × ~j). Thus in general, µ ~
R
and in particular, for currents, we have µ ~ ∝ L,
~ ~
~ which means that dJ ⊥ J~ and dJ ⊥ B, ~ or the same for µ
or more generally µ~ ∝ J, dt dt ~
~
instead of J, which means we get precession of µ ~
~ around B.
Quantum mechanically, that means that Lz , or Jz is constant, so in the electron
case, for given ms , e.g. of +1/2, so
|ψi = ψ(~r, t)|φ1 i. (17.14)
~ +S
6) For l = 2, since L ~ = J,
~ we have j = l − s, ..., l + s, in this case 3/2 or 5/2.
Then the states are
|3/2, mi , m = −3/2, −1/2, +1/2, +3/2 , and
|5/2, mi , m = −5/2, −3/2, −1/2, +1/2, +3/2, +5/2. (17.15)
7) To prove that
Sy
T = e−iπ ~ K0 , (17.16)
we note that K0 (iJi )K0−1 = −iJi , and the same for iJi replaced by iSi .
Sy Sy
But then e−iπ ~ Si e+iπ ~ for Si = Sy gives back Sy , for Si = Sz gives −Sz , and
for Sx gives −Sx .
18 Chapter 18
1) For the 3-body problem with 2-body potentials Vij (rij ), the Hamlitonian is
~2 ~ 2 ~2 ~ 2 ~2 ~ 2 X
Ĥ = − ∇1 − ∇2 − ∇ − Vij (rij ) , (18.1)
2m1 2m2 2m3 3 ij
~2 ~ 2 2
+ − (∇r31 − ∇ ~ r )2 − ~ (∇ ~r −∇ ~ r )2 − ~ (∇ ~r −∇ ~ r )2
23 23 12 12 31
2m3 2m2 2m1 i
×ψ̃(~r12 , ~r23 , ~r31 ) + (V12 (r12 ) + V13 (r13 ) + V23 (r23 )) ψ̃(~r12 , ~r23 , ~r31 ) φ(R)~
= (E − ECM )ψ̃(~r12 , ~r23 , ~r31 )φ(R) ~ , (18.7)
which splits into
~2 ~ ~ = ECM φ(R) ~ ,
− ∇R φ(R)
2M2 2 2
~ ~ ~ r )2 − ~ (∇ ~r −∇ ~ r )2 − ~ (∇ ~r −∇ ~ r )2
− (∇r31 − ∇ 23 23 12 12 31
2m3 2m2 2m1
+V12 (r12 ) + V13 (r13 ) + V23 (r23 )] ψ̃(~r12 , ~r23 , ~r31 ) = (E − ECM )ψ̃(~r12 , ~r23 , ~r31 ).
(18.8)
2) For the two-dimensional system with 1/r potential, V (r) = −Q̃2 /r, the re-
duced Schrödinger equation is
2
~ ~2
− ∇r + V (~r) ψ(~r) = Erel ψ(~r). (18.9)
2µ
In two dimensions, the Laplacian is
2 2
~ 2r = ∂ + 1 ∂ + 1 ∂ ,
∆r = ∇ (18.10)
∂r2 r ∂r r2 ∂φ2
2
∂
where ∂φ 2 = ∆φ is the Laplacian on the circle. The wave function separates vari-
ables as
ψ(r, φ) = R(r)eimφ , (18.11)
2
∂
such that ∆φ eimφ = ∂φ 2e
imφ
= −m2 eimφ . Then the resulting radial equation is
( )
~2 d2 m2 Q̃2
1 d
− + − 2 − R(r) = ER(r). (18.12)
2µ dr2 r dr r r
Defining the reduced radial function
χ(r)
R(r) = √ , (18.13)
r
the equation for χ(r) takes the form of the one-dimensional Schrödinger equation,
" !#
d2 χ(r) 1 − 4m2 2µ Q̃2
+ + 2 E+ χ(r) = 0 , (18.14)
dr2 4r2 ~ r
with the effective potential
Q̃2 ~2 m2 − 1/4
Veff (r) = − + . (18.15)
r 2µ r2
The boundary condition is χ(r = 0) = 0 at r = 0 and normalizability at infinity,
Z ∞ Z ∞
2
|ψ| rdr < ∞ ⇒ χ2 (r)dr < ∞. (18.16)
0 0
61 Chapter 18
so
µB ~ ~ ~ · B).
~
Ĥ 0 = Ĥ + (L · B + 2 S (18.48)
~
Then in the Hydrogen atom analysis, we shift
E → E − µB (m + 2ms )B. (18.49)
19 Chapter 19
1) For the potential V (r) = αr cos βr, Veff (r) is dominated at r → 0 by the term
l(l + 1)/r2 , so blows up. However, the potential (so also the effective potential) has
an infinite number of negative regions for r → ∞, so an infinite number of bound
states, like the Coulomb 1/r potential.
2) For the potential V (r) = |α|
r 2 ln r/r∗ , we have effectively s > 2 in the analysis in
the text, so the potential dominates over the l(l + 1)/r2 term at r → 0, so becomes
infinitely negative. As for the analysis in the text, we have states of arbitrarily
large |E| for E < 0, and the particle ”falls into r=0”, so the quantum particle is
effectively at r = 0 (in a bound state).
3) For the Yukawa potential V (r) = − |α| r e
−mr
, at r → 0, we have a Coulomb
potential, so again Veff is dominated by the l(l + 1)/r2 term, so we have a bound
state spectrum. However, at r → ∞, we have effectively s → ∞ in the analysis
in the text, so Veff is also dominated by the l(l + 1)/r2 at infinity, being therefore
positive. The effective potential therefore starts infinitely positive at r = 0, drops
down to a negative minimum, then up to a positive maximum, and then slowly
decays to zero. By the analysis of the energy of a state
~2 |α| −mr
E∼ − e >0 (19.1)
2m(∆r)2 r
for large r, there are no bound states at large distances, so we have necessarily a
finite number of bound states (at finite distance). From the shape of the potential,
we can indeed approximate it with a harmonic oscillator around the negative min-
imum (like for the diatomic molecule), and the approximation should be valid for
the first few excited states.
4) In the equality of the free wave functions (in Cartesian coordinates and in
spherical coordinates)
∞ X
l
~
X
eik·~r = alm (~k)jl (kr)Ylm (θ, φ) , (19.2)
l=0 m=−l
by choosing ~k onto Oz, since the left-hand side is independent of φ, the right-
imφ
hand side
q must also, so we must have m = 0 (the φ dependence is in e ). Since
2l+1
Yl,0 = 4π Pl (cos θ), with cos θ ≡ u and ρ = 2kr as usual, we have
∞
ρu X
ei 2 = ekr cos θ = cl jl (ρ/2)Pl (u). (19.3)
l=0
To find cl = (2l + 1)il c0 , we differentiate both sides of the above equation with
65
66 Chapter 19
On the other hand, using the recurrence relations of the spherical Bessel functions
l+1
jl0 (x) = jl−1 (x) − jl (x)
x
jl (x) 1
= (jl+1 (x) + jl−1 (x)) , (19.7)
x 2l + 1
we find
iux
X l l+1
iue = cl jl−1 (x) − jl+1 (x) Pl (u). (19.8)
2l + 1 2l + 1
l
Equating the two formulas for iueiux , and identifying the coefficients of Pl (u),
we find
cl cl1 cl cl+1
ljl−1 (x) −i = (l + 1)jl+1 (x) +i , (19.9)
2l + 1 2l − 1 2l + 1 2l + 3
but since this must be true at any x, we must put to zero the coefficient of jl−1 (x)
on the left and of jl+1 (x) on the right, giving the recurrence relation
cl+1 cl
=i , (19.10)
2l + 3 2l + 1
which iterated gives
cl = (2l + 1)il c0 . (19.11)
V̂ = V̂int (|X̂1 − X̂2 |) + V̂int (|X̂1 − X̂3 |) + V̂int (|X̂2 − X̂3 |) (20.2)
where
N (N − 1)
nN = (N − 1) + (N − 2) + ... + 1 = . (20.4)
2
4) The general state of the fermionic harmonic oscillator is
where b(~k) and c† (~k) must be different harmonic oscillators. This is so, since we
must have {ψ(~x, t), ψ † (~y , t)} = δ 3 (~x − ~y ), since ψ † is canonically conjugate to ψ
(the Lagrangian is linear in derivatives, not quadratic). Here we have
1) For the ortho-He atom with one electron in n = 1, l = 0, and the other in
n = 2, l = 0, so with spin singlet state, and symmetric spatial wave function,
1
φ = φs (~r1 , ~r2 ) = √ [ψq1 (~r1 )ψq2 (~r2 ) + ψq2 (~r1 )ψq1 (~r2 )] , (21.1)
2
in the approximation 1, the energy is just the sum of the energies,
e20 Z e2 Z
E = E (1) + E (2) = 2 + 0 2 , (21.2)
a 0 n1 a0 n2
so, for Z = 2 (He), we have
e2 5/2e2
2 2
E= 0 + 2 = . (21.3)
a0 12 2 (4π0 )a0
In the approximation 2, we need to add the electron-electron interaction, as
e2 e2
Z Z
∆E = hφs | 0 |φs i = d3 r1 d2 r2 0 |φs (~r1 , ~r2 )|2 = C + A , (21.4)
r12 r12
where
|ψn=1,l=0 (~r1 )|2 e2 |ψn=2,l=0 (~r2 )|2
Z Z
3
C= d r1 d3 r2
4π0 r12
Z Z 2
e ∗ ∗
A = d3 r1 d3 r2 ψn=1,l=0 (~r1 )ψn=1,l=0 (~r2 )ψn=2,l=0 (~r2 )ψn=2,l=0 (~r1 ).
4π0 r12
(21.5)
But
Rn0 (r) Ñn,0
ψn,l=0 (~r) = √ = √ e−kn r L1n−1 (2kn r) , (21.6)
4π 4π
and, since L10 (x) = 1 and L11 (x) = 2 − x and kn = Z/(na0 ), we find
3/2
2 2 2r
ψn=1,l=0 (~r) = √ e− a0
4π a0
3/2
2 1 − ar r
ψn−2,l=0 (~r) = √ e 0 1− . (21.7)
4π a0 a0
Then
2
27 e2
Z Z
4r1 2r2 r2
C= d3 r1 d3 r2 e− a0 e− a0 1 −
4πa60 4π0 |~r1 − ~r2 | a0
3(r1 +r2 )
7 Z 2 −
Z
2 3 3 e e a0
r1 r2
A= d r1 d r2 1− 1− . (21.8)
4πa60 4π0 |~r1 − ~r2 | a0 a0
71
72 Chapter 21
To calculate the integrals, we follow the example in the text. Indeed, there it was
shown that
e−4r1 a0
Z
Φ(~r2 ) ≡ d3 r1
r12
1 e−2k1 r2
1 −2k1 r2 1 1
= −π e + − , (21.9)
k12 k13 r2 k13 r2
for k1 = 2/a0 . Thus, we obtain (after some algebra) that
Z ∞ 2
27 e2
2
2r
− a2 r2
C= 4πr2 dr2 Φ(r2 )e 0 1−
4π(4π0 )a60 0 a0
2
2πe 13
= 1− 4 . (21.10)
a0 (4π0 ) 3
To calculate A, we need to calculate first
3r2
e− a0
Z
r2
Φ̃(r1 ) = d3 r2 1− , (21.11)
r12 a0
which (like for Φ(r1 ) in the text) satisfies (for k1 = 3r2 /a0 )
d2
r1 r1
∆r1 Φ̃(r1 ) = −4π 1 − e−2k1 r1 ⇒ 2 (r1 Φ̃(r1 )) = −4πr1 1 − e−2k1 r1 ,
a0 dr1 a0
(21.12)
so it has the particular solution of the type
But the solution to the homogeneous equation is r1 Φ̃(r1 ) = C1 r1 +C2 , and adding
the physical conditions Φ̃(r1 → ∞) → 0, that implies C1 = 0, and the condition
that Φ̃(r1 ) is nonsingular at r1 → 0, we find that the coefficient of 1/r1 in Φ̃ must
be 0, so C2 = −c.
Then
2
π r1 1 1 1 −2k1 r 1 1
Φ̃(r1 ) = + 2r1 −1 + − + e + − .
r1 k12 a0 a0 k1 k1 2a0 k12 k1 2a0 k12
(21.15)
Substituing k1 = 3/(2a0 ), we obtain
4πa20
2
r1 2 4 3r1 4
Φ̃(r1 ) = − r1 − a0 e− a0 + a0 . (21.16)
9r1 a0 3 9 9
Then, after some algebra, we find
Z ∞
2 7 e2
2
3r
− a1 r1
A= 4πr1 dr1 e 0 Φ̃(r1 ) 1 −
4π(4π0 )a60 0 a0
73 Chapter 21
24 · 95 e2
=− , (21.17)
37 a0 0
so finally
e2 1 24 · 95
13
∆E = C + A = − − . (21.18)
a0 0 2 2 · 34 37
2) For the He atom with two electrons in the n = 2, l = 0 state, the ground state
is para-He, since there is no antisymmetric spatial wave function (the antisymmetric
part is zero, because the electrons are in the same state).
Thus the wave function is
ψ = ψn=2,l=0 (~r1 )ψn=2,l=0 (~r2 )χsinglet
3
1 1 r +r
− 1a 2 r1 r2
= e 0 1− 1− . (21.19)
π a0 a0 a0
The energy in the first approximation is
Ze20 e2
E = E (1) + E (2) = 2 = , (21.20)
a 0 n2 (4π0 )a0
and in the second approximation, we add
e20
∆E = hψ| |ψi
r12
r1 +r2
e−2 a0
2 2
e20
Z Z
3 3 r1 r2
= 2 6 d r1 d r2 1− 1− . (21.21)
π a0 r12 a0 a0
Thus we must calculate
2r2
e− a0
Z 2
r2
Φ̃(r1 ) = d3 r2 1− . (21.22)
r12 a0
It satisfies
2
r1
∆r1 Φ̃(r1 ) = −4π 1 − e−2kr1 , (21.23)
a0
where k = 1/a0 . It becomes
2
d2
r1
r1 Φ̃(r1 ) = −4πr1 1 − e−2kr1 , (21.24)
dr12 a0
with the particular solution of the type
r1 Φ̃p (r1 ) = (ar13 + br12 + cr1 + d)e−2kr1 . (21.25)
From the differential equation, we fix the coefficients as (substituting also k =
1/a0 )
b π c 3π d π
a = −π , = , 2 = , 3 = . (21.26)
a0 2 a0 4 a0 8
To this, we must add a general solution to the homogenous equation, r1 Φ̃h (r1 ) =
C1 r1 + C2 and, as before, we must have Φ̃(r1 → ∞) → 0, implying C1 = 0, and
74 Chapter 21
= eff3 e− a0 2a0
. (21.32)
πa0 2
Then the energy in approximation 3 is
2
p~2 Ze2 Ze20 e2
p~1
E = hψs | + 2 |ψs i + hψs | − 0 − |ψs i + hψs | 0 |ψs i. (21.33)
2m 2m r1 r2 r12
But
p~2 Z 2 e20
h in = ⇒
2m 2a0 n2
p~2 p~2 p~2 p~2
1
hψs | 1 |ψs i = h 1 i1 + h 1 i2 + hψn=1,l=0 | 1 |ψn=2,l=0 ihψn=2,l=0 |ψn=1,l=0 i
2m 2 2m 2m 2m
2 2 2 2
1 Zeff e0 1 5 Zeff e0
= 1+ +0 = . (21.34)
2 2a0 4 16 a0
We also have
1 Z 1 1 1 1 1 5 Zeff
h in = ⇒ hψs | |ψ2 i = h i1 + h i2 = . (21.35)
r a0 n2 r1 2 r r 8 a0
75 Chapter 21
e2 e2 e2 e2
Z Z
C(R) = d3 r1d3 r2 − 0 − 0 + 0 + 0 |ψ(r1A )|2 |ψ(r2B )|2
r2A r1B r12 R
e20 e20 e20 e20
Z Z
3
A(R) = d r1 d r2 − 3
− + + ψ ∗ (r1A )ψ(r1B )ψ(r2A )ψ ∗ (r2B )
Z r2A r1B r 12 R
S = d3 rψ ∗ (rA )ψ(rB ) = hψrA |ψrB i ⇒ |S|2 < 1. (21.46)
We want to see that E+ (R) starts positive, decreases to a minimum below 0, then
goes up, asymptotically at infinity reaching 0 from below, while E− (R) decreases
until reaching 0 asymptotically from above.
The physical argument for that is that E+ corresponds to the symmetric spatial
wave function, which corresponds to the singlet spin function, ↑↓, which is stable,
so must have a minimum. On the other hand, E− corresponds to the antisymmetric
spatial wave function, thus to the triplet spin function, ↑↑, which is unstable, so
has no minimum. But at small R, C > A always, and the e20 /R term dominates, so
both start off positive. Then the only possibility is the one stated above.
6) Consider the 3 H atoms in a regular triangle. Why is it unstable? We have seen
that the stable H2 molecule is the one with spin singlet, ↑↓. But then by adding an
extra H, we will have ↑↑↓, which contains an unstable E− (R), so is unstable.
7) The Hamiltonian is
4 4 4 4
X ~2 X e20 X e20 X e20
Ĥ = − ∆i − + + . (21.47)
i=1
2m r
i,j=1 iAj
r
i,j=1 ij
R
i,j=1 Ai Aj
∂L ~ r
mc2 (m~v + q A)
pr = =q , (22.4)
∂vr ~ 2
(mc2 − qA0 )2 − c2 (m~v + q A)
so the Hamiltonian is
X (mc2 − qA0 )2 q ~
H= pr v r − L = q + A · p~. (22.5)
~ 2 m
r (mc2 − qA0 )2 − c2 (m~v + q A)
1 d d
Noting that √ = dτ , and and that we have defined
1−v 2 /c2 dt
muν + qAν
pν = q , (22.9)
~ 2
(mc2 − qA0 )2 − c2 (m~v + q A)
d dxµ
and that dt qAν = dt q∂µ Aν , we see that we would obtain the correct (relativistic)
Lorentz force law,
mduν
= q(∂ν Aµ − ∂µ Aν )muν , (22.10)
dτ
only if we neglect A0 and A ~ in the denominator of pµ . So in this case minimal
coupling almost works, but not quite.
We should note that the correct treatment is given by first noting that
r
2 2 dxµ dxν
Ldt = −mc dτ = −mc − ηµν dτ , (22.11)
dτ dτ
and then adding the coupling to electromagnetism as the term Aµ dxµ pulled back
R
(i~∂t )2 ψ = Ĥ 2 ψ , (22.15)
~ · dS
~ 6= 0 through a surface S
R
3) If we have a nonzero magnetic flux, Φ = S
B
bounded by the conductor closed loop C, then
Z I
~ ~
Φ = (∇ × A) · dS =~ ~ · d~l 6= 0.
A (22.18)
S ∂S=C
79 Chapter 22
and, in ~j = ~jn + ~js , we neglect the normal current ~jn , since it vanishes quickly in
the absence of external fields. For the same reason, we put E ~ = 0. Then ~j = ens~vs
(the superconducting current is composed of quasiparticles with charge e, effective
mass m∗ , density ns and velocity ~vs ), and the kinetic energy associated with it is
Z
~v 2
Z ~j 2 Z
µ0 λ2L~js2
f1 = ns m∗ s = m∗ 2s = , (22.23)
2 2e ns 2
where we have introduced, as in the text, µ0 λ2L ≡ m∗ /(ns e2 ). Then we obtain the
formula we needed for the total free energy in (22.21), from f = f0 + f1 .
Minimizing (22.21) over ~j, δf /δ~j = 0, we obtain
!
Z
δ ~
B
B~· + λ2L µ20~j = 0. (22.24)
δ~j
~
But, from ~js = −A/(µ 2 ~
0 λL ), taking ∇×, we obtain
~ = −µ0 λ2L ∇
B ~ × ~js . (22.25)
~ ~j =
Then δ B/δ ~
−µ0 λ2L ∇× δ(~x − ~y ) under the integral sign, and by partial inte-
~ × B,
gration in (22.24), we get ∇ ~ more precisely
~ ×B
+µ0 λ2L ∇ ~ + ~j = 0. (22.26)
~ of the above, then using again ∇
Taking ∇× ~ × ~j = − 1 2 B,
~ and that ∇
~ × (∇
~ ×
µ0 λL
~ =∇
B) ~ 2 B,
~ we finally get
~ 2B
λ2L ∇ ~ = B.
~ (22.27)
5) When the magnet Rgoes down through the middle of the superconducting ring,
the magnetic flux Φ = S B ~ · dS
~ increases as well. But since B~ = −µ0 λ2 ~js inside
L
the superconductor, we get that
Z Z I
~ ~ 2 ~ ~ ~
B · dS = −µ0 λL (∇ × js ) · dS = −µ0 λL 2 ~js · d~l , (22.28)
S S C=∂S
80 Chapter 22
However, since B ~ acts in opposite directions to dS ~ in the two loops of the figure
eight, if the size of the loops is identical, the two contributions to the magnetic flux
cancel, so there is no Aharonov-Bohm phase.
2) As we undo the figure eight into a circle, the total magnetic flux increases
from 0 to nonzero, and so does the Aharonov-Bohm phase, such that in the end
iq R
~ ~l
A·d
|ψ(t)i = e ~ C |ψi. (23.3)
~ i , i = 1, 2, ..., N , the generic Berry phase is a sum,
3) For positions R
Z
1X ~ (i) · dR
~i ,
γn = A n (23.4)
~ i
81
82 Chapter 23
where
∂ ~ − ∂ Ai (E).
~
Fijn = ∂i Anj − ∂j Ani = Aj (E) (23.12)
∂Ei ∂Ej
∂
Since ∂E i
Ĥint = di , if we assume that there are eigenfunctions in the directions
~ called Oz, we obtain
of E,
di
Bin = . (23.13)
E2
Then, as in the case in the text, the Berry phase is
I Z Z Z ˆ
di 2 ~ di 2
γn = Ani dEi = Bin d2 Ei = − 2
d E i = −|d| 2
d Ei
C=∂S S S E E
~
= |d|∆Ω cos(E, ~ ,
~ d) (23.14)
where E ~ is in the center of the (small) S.~ Unlike the case in the text, we see the
appearence of the cosine, so the result is quasi-geometric.
B̂i
6) In order to prove that Bin = m~ B 2 in the case in the text, first consider the
then the second-order perturbation theory formula (23.54) in the text adapted for
our case,
" #
X hn(B(t))|∂i Ĥ|m(B(t))i
n
Fij = i hm|∂j Ĥ|ni − (j ↔ i) . (23.16)
(Em − En )2
m6=n
Since
gµB g
∂i Ĥ = − Si = − µB σi , (23.17)
~ 2
If the spin S~ is aligned with the z direction (which we can always choose), then
|mi is an eigenstate of Sz , so hm|∂z H|ni ∝ hm|ni = 0 (since m 6= n). Therefore,
we obtain Fiz = 0, as we should. Moreover, the only nonzero component of Fij is
Fxy , so (using that Em − En = 2 g2 µB B, and the explicit forms of the σx = σ1 and
σy = σ2 Pauli matrices), for |ni = | + 1/2i fixed we have
i
Bzn = Fxy
n
= [h+1/2|σx | − 1/2ih−1/2|σy | + 1/2i − (x ↔ y)]
4B 2
1
= − 2. (23.18)
2B
83 Chapter 23
But there is a possibility to have also |ni = | − 1/2i, in which case we obtain
Bz = +1/B 2 . Al, in all, we have the formula we wanted,
B̂i
Bin = −m , (23.19)
B2
with m = ±1/2.
7) In the covariant non-Abelian case, we can define the non-Abelian Berry con-
nection associated with states |mi of the same energy (degenerate) as
(nm)
Ai (~k) = ihn, ~k(t)|∇
~ k |m, ~k(t)i ,
i
(23.20)
and in the usual way the field strength of the non-Abelian connection,
(nm) (nm) (nm) (kl) (qr)
Fij (~k) = ∂i Aj − ∂j Ai + f (nm) (kl)(qr) Ai Aj , (23.21)
in terms of the structure constants f with inidices (mn) in the adjoint of U (N ),
leading to the usual F = dA + A ∧ A.
24 Chapter 24
h/e
∆B = . (24.3)
Lx Ly
where
~ B)
A( ~ = hn(B(t))|
~ ~ ~ |n(B(t))i.
∇ ~ (24.11)
B
Z Z
B̂i 2
= −mJ d Bi = −mJ dΩ , (24.15)
S(C) B2 C
just like in the spin 1/2 case. We note that the formula is actually independent on
g!
It remains to prove (24.14). As before, we write (substituting ∂i Ĥ and Em in the
general formula)
" Ŝ
#
n i X hn(B(t))| Ŝ~i |m(B(t))ihm(B(t))| ~j |n(B(t))i
Fij = 2 − (↔ j) . (24.16)
B ∆m2J
m6=n
where we used
gµB
∂i Ĥ =Ŝi , Em − En = ∆mJ gµB B , (24.17)
~
and we note that gµB cancels between the numerator and denominator.
Then we note that Fzi = 0, where B = Bz , and that
" Ŝy
#
Ŝx
i X hn J | |m J ihm J | |nJ i
Bzn = Fxy
n
= 2 ~ ~
− (x ↔ y) , (24.18)
B (mJ − nj )2
mJ 6=nJ
Ŝx 1
hnJ | |mJ i = [αJnJ δnJ ,mJ +1 + αJ,nJ +1 δnJ ,mJ −1 ]
~ 2
Ŝy 1
hnJ | |mJ i = [αJnJ δnJ ,mJ +1 − αJ,nJ +1 δnJ ,mJ −1 ]
~ 2i
p
αJ,nJ = (J + nJ )(J − nJ + 1) , (24.19)
we find
"
n i Ŝx Ŝy Ŝx Ŝy
Fxy= 2 hnJ | |nJ + 1ihnJ + 1| |nJ i − hnJ + 1| |nJ ihnJ | |nJ + 1i
B ~ ~ ~ ~
#
Ŝx Ŝy Ŝx Ŝy
+hnJ | |nJ − 1ihnJ − 1| |nJ i − hnJ − 1| |nJ ihnJ | nJ − 1i
~ ~ ~ ~
1 1
= (α2 2
− αJ,n ) = − 2 nJ . (24.20)
2B 2 J,nJ +1 J
B
Then finally, indeed
B̂i
Bin = nJ . (24.21)
B2
25 Chapter 25
and find
s001 + (s01 )2
s02 = −
2s00 ( )
1 1 p 00 1 p 0 2
= −p − ln 2m(E − V (x)) + ln 2m(E − V (x)) .
2m(E − V (x)) 2 4
(25.4)
2
2) For a harmonic oscillator in 1 dimension, V (x) = k x2 , we have two turning
points x1 , x2 , in the first the potential is going down, and in the second, it is going
up. So we can use the formulas in the text for the turning point identifications, at
x1
" s #
1 x1 0
Z 2
1 x
x2 1/4 exp − ~ dx 2m k
2
−E
k 2 −E x
√ " Z r #
2 1 x 0 x2 π
→ 2 1/4
sin dx 2m(E − k ) + , (25.5)
E − k x2 ~ x1 2 4
87
88 Chapter 25
√ " Z r #
2 1 x2 0 x2 π
→ 2 1/4
sin dx 2m(E − k ) − . (25.6)
E − k x2 ~ x 2 4
3) Assuming a wave function centered on one of the minima, then x1 and x2 are
turning points on each side of the minimum, so at x1
1 x1 0 p
Z
1
1/4
exp − dx 2m (V0 cos ax − E)
[V0 cos ax −
√ E]
~ x
Z x
2 1 0
p π
→ 1/4
sin dx 2m(E − V0 cos ax) + , (25.7)
[E − V0 cos ax] ~ x1 4
0 3|α| 2µ
Veff (r0 ) = 0 ⇒ r0 = , (25.10)
2l(l + 1) ~2
and the value at this maximum is
4 ~2 l(l + 1)
Veff (r0 ) = . (25.11)
27 2µ |α|2
Define
p
k(r) = p2m(E − Veff (r)) (region III)
k̃(r) = p2m(E − Veff (r)) (region I)
κ(r) = 2m(Veff (r) − E) region II). (25.12)
Then at r > r2 (region III), we have
1 i Rrr dr0 k(r0 ) −i r dr 0 k(r 0 )
R
ψIII (r) = p e 2 + e r2 , (25.13)
k(r)
and the turning point formula at r2 means that we have in region II (r1 < r < r2 )
Z r2
1 τ 0 0
ψII (r) = p e exp − dr κ(r ) , (25.14)
κ(r) r
where
Z r2
τ= dr0 κ(r0 ). (25.15)
r1
89 Chapter 25
Then the turning point formulas at r1 means that we have in region I (r < r1 )
A i R r1 dr0 k̃(r0 )+β Rr 0 0
+ e−i r dr k̃(r )+β .
1
ψI (r) = q e r (25.16)
k̃(r)
Note that these formulas are the same ones as in the text, in the case of a well
potential.
5) Matching ψ and ψ 0 at r = r1 now gives us A = eτ .
The transmission coefficient is
|jtrans |
T = , (25.17)
|jinc |
where
~j ≡ ~ ψ ∗ ∇ψ
~ − ψ ∇ψ
~ ∗ . (25.18)
2mi
Then
1) We have
q̇ 4 q̇ 2 (q̇ 2 + α)2 α2
L= + α − λq 4 = − − λq 4 . (26.1)
4 2 2 4
Then the conjugate momentum is
∂L
p= = q̇(q̇ 2 + α) , (26.2)
∂ q̇
and the quantization condition is
I I
1 1
pdq = n + h⇒ q̇(q̇ 2 + α)q̇dt = n + h. (26.3)
C 2 C 2
2) For the potential V = k|x|, at x → ±∞, we have
1 x 0p
Z
1
ψIII,W KB = exp − dx 2m(k|x| − E)
[2m(k|x| − E)]1/4 ~ x2
const. 4mk 3/2 2mE 1/2
→ exp − x − x . (26.4)
(2mk|x|)1/4 3~ ~
On the other hand, at x → 0, we have
Z x
1 1 0
p π
ψII,W KB (x) = cos dx 2m(E − k|x|) −
[2m(E − k|x|)]1/4
~ x1
r 4
1 2mEn k m 2
→ cos x− x +δ . (26.5)
(2mEn )1/4 ~ 2E 2En
3) For a particle in a box, −L/2 ≤ x ≤ L/2, the Bohr-Sommerfeld quantization
gives
Z +L/2 √ √
0 1
dx 2mE = 2mEL = n + π~ , (26.6)
−L/2 2
so
2
π 2 ~2
1
En = n+ , (26.7)
2 2mL2
so we only need to replace n + 1/2 with n in order to get the exact result. That
means that at large n, we get the correct result.
Using this En , the WKB wave function becomes
" Z #
1 1 L/2 0 p π
ψ(II)W KB (x) = cos dx 2mEn −
(2mEn )1/4 ~ −L/2 4
90
91 Chapter 26
s
L 1 πx nπ
= cos n+ + . (26.8)
π~(n + 1/2) 2 L 2
Then for n odd, we get ∓ sin[(n + 1/2)πx/L], and for n even, we get ∓ cos[(n +
1/2)πx/L].
That means that the wave function is OK, except for the normalization constant,
and an overall phase in the cosine.
4) For the harmonic oscillator in the Bohr-Sommerfeld quantization, in the clas-
sical n → ∞ limit, we get
mω 2 x2
ψn (x) ' const.xn e− ~ 2 , (26.9)
and then the wave function is approximately exact in the large x limit only. But
then, large x is allowed (classical oscillators reach there), and xn can be of the
mω 2 x2
same order of magnitude as e ~ 2 , so |ψn (x)|2 ∼ 1 (large probability there, since
a classical oscillator can reach there).
5) To write the equations for R2 (r) and Θ2 (θ), we need to keep the next order
in (26.29). This is
2 " 2 2
~ ∂s1 ∂s0 ∂s2 1 ∂s1
∆s1 + +2 + 2
i ∂r ∂r ∂r r ∂θ
2 #
2 ∂s0 ∂s2 1 ∂s1 2 ∂s0 ∂s2
+ 2 + 2 2 + 2 2 = 0.
r ∂θ ∂θ r sin θ ∂φ r sin θ ∂φ ∂φ
(26.10)
The separated variable ansatz is
s0 = R0 (r) + Θ0 (θ) + Φ0 (φ)
s1 = R1 (r) + Θ1 (θ) + Φ1 (φ)
s2 = R2 (r) + Θ2 (θ) + Φ2 (φ). (26.11)
But ∂s0 /∂φ = ∂Φ0 /∂φ = L3 and then d2 Φ0 /dφ2 = 0, and in turn, it means we
can choose ∂s1 /∂φ = ∂Φ1 /∂φ = 0, since we need d2 Φ1 /dφ2 = 0. Then, since
d2 R0 1 d2 Θ0
2 dR0 dΘ0
∆s0 = + + 2 + cot θ
dr2 r dr r dθ2 dθ
2 2
d R1 2 dR1 1 d Θ1 dΘ1
∆s1 = + + 2 + cot θ , (26.12)
dr2 r dr r dθ2 dθ
we find the (separated) equations
2
d2 R1 (r) 2 dR(r)
dR1 dR0 dR2
+ + + =0
dr2 r dr dr dr dr
2 2
1 d Θ1 dΘ1 1 dΘ1 2 dΘ0 dΘ2
+ cot θ + 2 + =0
r2 dθ2 dθ r dθ r dθ dθ
2L3 dΦ2
=0, (26.13)
r2 sin2 θ dφ
so again we can choose Φ2 (φ) = 0.
92 Chapter 26
6) For the central potential V = −α/r2 , there is only one positive solution to
the vanishing of the E − Veff (r),
2mE − L2
2mE + =0, (26.14)
r2
(and only for 2mE < L2 do we have even one solution), since the effective potential
starts at +∞ and decays monotonically to zero.
That means that there is only one condition imposed on the wave function (at
infinity), not two (at zero and infinity), so there is no quantization, all positive
energies are possible!
7) For the hydrogenoid atom, the radial wave function is
" r #
1 r 2mZe20 L2
Z
1
Rn (r) = exp − dr 2m|E| + − 2
[2m(E − Veff (r))]1/4 ~ r2 r r
√
" p #
2
const. m Ze 0 2m|En |
' r− 2En ~ exp − r . (26.15)
(2m|En |)1/4 ~
1) The answer is yes, if there are both electric and magnetic sources, i.e., if the
action is Z Z
4 1
S = − d x Fµν F + d4 x(j µ Aµ + k µ õ ) ,
µν
(27.1)
4
where j µ is the electric current, k µ is the magnetic current, and õ is the dual
gauge field, such that the dual field strength is F̃µν = ∂µ õ − ∂ν õ . Then the
Maxwell duality is
Aµ ↔ õ , j µ ↔ k µ , Fµν → ∗Fµν = F̃µν . (27.2)
For a set of particles minimally coupled to electromagnetism, their electric/magnetic
charges must also be dual, Qe ↔ Qm . But then, in the quantum theory, this is only
possible if the particle spectrum is also invariant.
2) For an infinitesimally thin, planar, infinite perfect conductor, we will have
that the parallel electric field at the surface of the conductor, E ~ || = 0, and the
transverse magnetic field at the surface of the conductor, B ~ ⊥ = 0.
Indeed, if the conductor is perfect (σ = ∞), inside it E ~ =E ~ || = 0 (since otherwise
~j = ∞, which is impossible), and B ~ ⊥ = 0 (since otherwise, electrons would feel a
parallel Lorentz force that would have nothing to balance it, so the same would
happen).
But then, we can use the Maxwell equations to prove the continuity conditions
at a surface, applied here to the air/conductor surface, which are ~n × ∆E ~ = 0,
~n · ∆B = 0 (so no change in E ~ || and B ~ ⊥ at the interface), as well as ~n × ∆H ~ =
~j, ~n · ∆E = σ/0 .
The first relation leads to the fact that E ~ || = 0 just outside of the surface of
the conductor, which leads to the familiar fact that an electric charge put near the
conductor will generate a mirror charge (in reality, the charges in the conductor
rearrange to create this illusion) of same value and opposite sign (Q0e = −Qe ),
situated at an equal distance from the conductor, on the other side.
To prove the second relation, one considers a rectangular contour with two lines,
one inside the conductor, one outside, both parallel to the interface, and closed by
two very small lines perpendicular to the surface (which can then be neglected).
Then the Maxwell equation ∇ ~ ×E ~ = − ∂ B, ~ integrated over the contour’s surface
∂t
gives
Z I
~ ~ ~ ∂ ~ ~
0 = (∇ × E)dS = − B · dl. (27.3)
S C=∂S ∂t
~ is parallel to the surface S (which is transverse
The left-hand side is zero since E
93
94 Chapter 27
to the surface of the conductor), due to the previous relation. Then we get that
∂ ~
∆ ∂t B = 0 on the two sides of C (inside and outside the conductor), which really
means (for a static situation) ∆B ~ ⊥ = 0. But since we already knew that B ~⊥ = 0
~
inside the conductor, it follows that B⊥ = 0 just outside it, too.
Then in the presence of the conductor the magnetic charge gets also a mirror
charge, just of the same sign (Q0m = +Qm ).
So, for an electric charge e and a magnetic charge g situated at the distance R
from the conducting plane, there are mirror charges −e and +g on the other side,
at the same distance R, leading to
~ =E
~ ⊥ = 2E
~1 = 2e ~ =B
~1 − B
~1 = B
~ || = 0.
E ~e⊥ , B (27.4)
4π0 R2
3) Consider the two particles (dyons, of identical masses m) with vectors ~r1 and
~r2 . Then the time derivative of the total angular momentum is
~
dL
= m(~r1 × ~r¨1 + ~r2 × ~r¨2 ). (27.5)
dt
They each feel Lorentz forces due to the other particle, so
m~r¨1 = q1 (E
~2 + ∇
~ ×B
~ 2) , m~r¨2 = q2 (E
~1 + ∇
~ ×B
~ 1 ). (27.6)
~ = ~r1 − ~r2 (so is the vector from particle 2 to particle 1), and choose a
Define R
reference frame where (instantaneously, at the moment t) ~v2 = 0. Then for particle
2 we can use the usual formulas,
~ 2 = q2 R̂ , B
E ~ 2 = µ0 g2 R . (27.7)
4π0 R 2 4π R2
For particle 1, however, which now has a generically nonzero velocity ~v1 , we must
use the Lorentz transformation rules,
~0 = E
E ~ || E
~⊥0 ~⊥ − β
= γ(E ~ × B)
~
||
~ =B
B 0 ~ || B
~ = γ(B
0 ~ ⊥ + β~ × E).
~ (27.8)
|| ⊥
Then
m~r¨2 = q2 (E
~ 10 + ~v2 × B
~ 20 ) = q2 (E
~ 1 − ~v1 × B
~ 2 ). (27.10)
Finally then
~
dL
= q1 (~r1 × E~ 2 + ~r1 × (~v1 × B
~ 2 )) + q2 (~r2 × E~ 1 − ~r2 × (~v1 × B~ 1 ))
dt
q1 q2 ~ + ~r1 × (~v1 × q1 B~ 2 ) − ~r2 × (~v1 × q2 B
~ 1)
= (~r1 − ~r2 ) × R
4π0
95 Chapter 27
!
µ0 q1 g2 R̂ µ0 q2 g1 R̂
= −~v1 × ~r1 × 2
+ ~r2 × . (27.11)
4π R 4π R2
Now choosing (instantaneously, at time t) ~r2 = −~r1 (so the origin to be at the
midpoint between the two particles), so that R2 = 4r12 , we have
!
~
dL µ0 R̂
= (q1 g2 − q2 g1 )~r1 × ~v1 × 2 . (27.12)
dt 4π R
~ tot must be an
and by the same argument as in the text, the extra term added to L
integer N times ~/2, which gives
q1 g2 − q2 g1 = 4π~N , (27.16)
which is the DSZ quantization condition, except for a factor of 2 (we only get it for
even N ).
4) We have eg2 − q2 h/e = h, so g2 = h/e(1 + q2 /e). Assume particle 2 is at rest,
then
~ 2 ) = q1 q2 R̂ + ~v1 × µ0 q1 g2 R̂
~ 2 + ~v1 × B
mr̈1 = q1 (E
4π0 R2 !4π R
2
e h R̂ µ0 h q2 R̂
= + ~v1 × 1+ . (27.17)
4π e0 R2 e e R2
~
5) For B(x) = 2θ 2
e δ (~
x),
the Berry phase is
I I
γ=e ~ ˙~
Adl = ~ · dS
B ~ = 2θ. (27.18)
C=∂S S
Then eiγ = e2iθ , but rotating one particle around the other amounts to a double
interchange of the particles. That means that we get the phase eiθ for interchanging
of identical particles. This is the definition of an ”anyon”, a particle with a fractional
statistics (a phase different than ±1 at interchanging identical particles).
6) The sphere has the monopole and antimonopole on it, at opposite sides. Then
96 Chapter 27
the total magnetic charge inside the sphere is zero, even if we add (part of) the m
and m̄ charges. But then the magnetic flux through the sphere is zero, giving
I I
0= (∇~ × A)
~ · dS
~= A~ · d~l , (27.19)
S2 ∂S 2
which contains no contradictions (the integral is zero, since the sphere has no bound-
ary). So a unique A~ is well defined on the sphere.
7) Yes, we can still infer that the electron charges are quantized. The Dirac
quantization can be defined by a surface going in between the monopole and anti-
monopole, so it is still valid.
28 Chapter 28
poles inside the closed contour, so the result is zero, ∆(t, t0 ) = 0. Therefore, all in
all, we have
0 0 1 −iω(t−t0 ) 1 +iω(t−t0 )
∆(t, t ) = θ(t − t ) + e − e . (28.7)
2ω 2ω
In this case, the boundary conditions are: at t → +∞, we get
e−iωt − eiωt
q(t) ∼ ∆ ∼ , (28.8)
2
whereas -at t → −∞, we get
q(t) ∼ ∆ ∼ 0. (28.9)
the Wick rotation back to Minkowski space is t = −itE , so EE = −iE, but in fact
we must rotate with only π/2 − angle, so EE → −i(E + i0 ), and
2
EE + ω 2 → −EE2
+ ω 2 − i ⇒
(E1,E + E2,E − E)2 + ω32 → −(E1 + E2 − E)2 + ω32 − i , (28.16)
where d2 θ = − 14 dθα dθβ αβ , we need to remember that dθ1 = 0 and θ12 =
R R R
θ22 = 0 = (θ2 )2 , which means that we will only select the terms with two thetas,
specifically with θ1 θ2 , from the integrand. Moreover, note that
Z Z
1
d2 θθ2 = − dθ1 dθ2 (2θ2 θ1 ) = +1. (28.22)
2
Then
Z
d2 θa1 Φ = a1 F (x) ,
Z Z
d2 θa2 Φ2 = a2 [2φ2 F (x) + 2 d2 θ(θα ψα (x))2 ]
Z Z
= a2 [2φ(x)F (x) − dθ1 dθ2 ((θ1 ψ1 )2 + (θ2 ψ2 )2 + 2θ1 ψ1 θ2 ψ2 )]
7) We divide the interval from t to t0 into N + 1 steps, and use the completeness
100 Chapter 28
relations Z
dψi dψ̄i |ψi ihψ̄i |e−ψ̄i ψi = 1l , (28.24)
q.e.d.
29 Chapter 29
φ1 = p2 − αq1 ≈ 0. (29.16)
The Hamiltonian is
p21
H = p1 q̇1 + p2 q̇2 − L = + βq1 q22 . (29.17)
2
The (first) secondary constraint is
and gives
βq2 (2p1 + αq2 )
βq2 (2p1 + αq2 ) + U (−α2 − 2βq1 ) = 0 ⇒ U = . (29.21)
α2 + 2βq1
There is no solution to the homogenous equation, since there is a single constraint
φ1 , and the equation would be
{f, g}D.B. = {f, g}P.B. − {f, χs }P.B. css0 {χs0 , g}P.B. , (29.23)
where
−1
css0 = ({χs , χs0 }P.B. ) , (29.24)
c̃12 = {φ1 , φ2 }P.B. = {p2 − αq1 , −αp1 − 2βq1 q2 }P.B. = α2 + 2βq1 ≡ A. (29.25)
0 A 0 −1/A
The matrix is , with inverse matrix , so
−A 0 1/A 0
1
c12 = − . (29.26)
A
Then
1 1
{f, g}D.B. = {f, g}P.B. + {f, φ1 }P.B. {φ2 , g}P.B. − {f, φ2 }P.B. {χ1 , g}P.B. .
A A
(29.27)
5) For
s
2
F µν
Z
Fµν 1 µνρσ
LB.I. = −L−4 d4 x 1 + L4 − L8 Fµν Fρσ − 1 , (29.28)
2 8
P0 = 0 , (29.30)
The Hamiltonian is
Z
H = d3 xPµ Aµ,0 − L
Z 1
= d3 x q ∂0 Ai F i0
4 Fµν F µν
8 1 µνρσ
2
1+L − L 8 Fµν Fρσ
2
L4 µ0 ν 0 ρ0 σ0
+ Fµ0 ν 0 Fρ0 σ0 40µνρ ∂0 Aµ Fνρ + L−4
2 · 64
L4 µνρσ
1 µν 2
+ Fµν F − ( Fµν Fρσ ) − 1
2 64
104 Chapter 29
Z
1 1 0i 0i 1
= d x 3
q F F + Fij F ij + L−4
4 F µν F µν
8 1 µνρσ
2 2 2
1+L 2 − L 8 Fµν Fρσ
L4
2
− (µνρσ Fµν Fρσ ) − L−4 + P i ∂i A0
2 · 64
−4 1 ij
L + 2 Fij F
Z 1
−4
= d3 x P i F i0 − A0 ∂i P i + q − L ,
2 F F µν 2
1 + L4 µν2 − L8 81 µνρσ Fµν Fρσ
(29.32)
so, since all the constraints are first class, the Dirac brackets are equal to the Poisson
brackets.
Then the primary constraints are φm : φq = P 0 , and the secondary constraints
are φj : φ2 = P i ,i . Then as in the Maxwell case,
Z
H 0 = H + d3 xU P 0 (~x)
Z
HT = H + d3 xv(~x)P 0 (~x0 )
0
Z Z
= d x{...} + d3 x[−A0 P i ,i + vP 0 ] ,
3
(29.37)
Then we obtain
{ψ, p}P.B. = 1 ⇒ {ψ, p}D.B. = 1 , {ψ ∗ , p∗ }P.B. = 1 ⇒ {ψ ∗ , p∗ }D.B. = 0.
(29.52)
It means that using Dirac brackets, we can put p∗ = 0 and p = −iψ ∗ . But still
HT = 0.
30 Chapter 30
1) The state is
1
|χa1,2 i = √ |ψ1,2 i + a|φ1,2 i
2
1
= √ √ [| ↑A ↑B i ± | ↓A ↓B i + a| ↑A ↓B i ± a| ↓A ↓B i] . (30.1)
1 + a2 2
But the most general separable state is
c (| ↑A i + α| ↓A i) ⊗ (| ↑B i + β| ↓B i)
= c [| ↑A ↑B i + αβ| ↓A ↓B i + α| ↓A ↑B i + β| ↑A ↓B i] . (30.2)
In order for this to equal the previous, we must have
{ α = ±β = a , αβ = ±1 } ⇒ ±β 2 = ±1. (30.3)
That means that β 2 = 1, so β = ±1, so a = ±1. For a = +1, we have
1 1
√ (|ψ1 i + |φ1 i) , √ (|ψ2 i + |φ2 i) , (30.4)
2 2
whereas for a = −1, we have
1 1
√ (|ψ1 i − |φ1 i) , √ (|ψ2 i − |φ2 i) . (30.5)
2 2
Equivalently,
1
ρA = TrB [| ↑A ↑B i ±1 | ↓A ↓B i + a| ↑A ↓B i ±2 | ↓A ↑B i]
2(1 + a2 )
⊗ [h↑A ↑B | ±1 h↓A ↓B | + ah↑A ↓B | ±2 h↓A ↑B |]
1
(1 + a2 )| ↑A ih↑A | + (1 + a2 )| ↓A ih↓A i
= 2
2(1 + a )
+a(±1 ±2 )| ↓A ih↑A | + a(±1 ±2 )| ↑A ih↓A |] . (30.6)
That means that we need a(±1 ±2 ) = ±(1 + a2 ), which is indeed what we had
found.
2) We have
1 1
|χa1,2 i = | ↑iA ⊗ p (| ↑i + a| ↓i)B +| ↓iA ⊗ p (±1 | ↓i ±2 a| ↑i)B ,
2(1 + a2 ) 2(1 + a2 )
(30.7)
and we can define the states after the tensor products as |ĩiB . If ±1 ±2 = 0, then
we find
1
ρA = (| ↑iA | ↑iB + | ↓iA | ↓iA ) , (30.8)
2
107
108 Chapter 30
√
so pi = 1/2, meaning that |i0 iB = 2|ĩi. Then finally
1
|10 i = √ (| ↑i + a| ↓i)B
1 + a2
1
|20 i = √ ± (| ↓i − a| ↑i)B ,
1 + a2
1
|χa1,2 i = √ (| ↑iA ⊗ |10 iB + | ↓iA ⊗ |20 iB ) . (30.9)
2
3) Consider the unitary matrix UAB defined in eq. (30.26) in the text. Then
| ↑↑i |ψ1 i | ↑↓i |φ1 i
UAB = , UAB = ⇒
| ↓↓i |ψ2 i | ↓↑i |φ2 i
a
1 | ↑↑i + a| ↑↓i |χ1,2 i
UAB √ = , (30.10)
1 + a2 | ↓↓i + a| ↓↑i |χa1,2 i
−1
so UAB disentangles them.
4) We have
so
S = − Tr ρ̂ ln ρ̂ = −λ1 ln λ1 − λ2 ln λ2 . (30.15)
5) We have
we obtain
P (~a−, ~c+) ≤ P (~b+, ~c+) + P (~a−, ~b+) , (31.10)
which in quantum mechanics is
" #
1 ~ c) a, ~b)
2 θ(~
a, ~c) 1 2 θ(b, ~ 2 θ(~
cos ≤ sin + cos . (31.11)
2 2 2 2 2
so for instance
|[(A + A0 ) + (A + A00 ) + (A0 − A00 )]B + [(A − A0 ) + (A − A00 ) − (A0 − A00 )]B 0 | ≤ 6 ⇒
|2hABi + 2hA0 Bi + 2hAB 0 i − 2hA0 B 0 i − 2hA00 Bi| ≤ 6 ⇒
|C(a, b) + C(a0 , b) + C(a, b0 ) − C(a0 , b0 ) − C(a00 , b)| ≤ 3. (31.21)
To violate this in quantum mechanics, consider that a, b, a0 , b0 are at small angles
from each other and a00 is at π from ~b. Then the violation is as before.
6) We can consider as an ”experiment” the transition from quantum to classical
at a later point. That ”measures” the wave function of the Universe. We obtain
probabilities and compare to the ”experiment” (our classical Universe).
7) We have very small deviations, at each moment in time. Then the probabil-
ity that the (genetically the) same person is born, let alone that it has the same
characteristics (personality, etc.), is ludicrously small.
32 Chapter 32
S = kB ln Ω = N kB ln 2. (32.5)
X P N
Y
Z= e−β i (Eni −µi Ni ) = Zi
{ni } i=1
∞ ~ωi
X
−β(Ei −µni ) e−β 2
Zi = e = . (32.15)
ni =0
1 − e−β(~ωi −µ)
To see that, note that at large n, there are ' np(x) letters ax , out of n. Then
n!
the combinatorics of groups of npx ax ’s gives Q (np x )!
, since (np(x))! is the number
x
of permutations of these objects. So the combinatorics is the number of ways of
picking these letters = total number of permutations, divided by the permutations
of each object. √
√ Then, the Stirling approximation gives N ! ' 2πN N −1/2 e−N , and we neglect
2πe−N . Then we note that 2nH(x) = x p(x)−np(x) , and by substituting the
Q
where S(ρi ) = − Tr(ρi log ρi ), consider first only ρ1 and ρ2 . Then, for t ∈ [0, 1],
ρ(t) = tρ1 + (1 − t)ρ2 , S(ρ(t)) = − Tr(ρ(t) log ρ(t)) (33.5)
implies
dS ρ̇(t)
= − Tr[ρ̇(t) log ρ(t)] − Tr ρ(t) . (33.6)
dt ρ(t)
But Tr ρ(t) = 1, so Tr[ρ̇(t)] = 0, so we are left with the first term, in which we
note that ρ̇(t) = ρ1 − ρ2 is independent on t. In order to take a second derivative
of t, that will then act only on ρ(t), we must re-express it as
Z ∞ Z ∞
1 1 d 1 1
log ρ(t) = ds − ⇒ log ρ(t) = ds ρ̇(t) .
0 s + 1 s + ρ(t) dt 0 s + ρ(t) s + ρ(t)
(33.7)
117
118 Chapter 33
d2 S
≤0, (33.9)
dt2
so the von Neumann entropy is concave.
4) For an example of S(A + B) < S(A) + S(B), consider the case of A + B in
the pure state |ψi that is entangled, e.g.
1
|ψi = √ (|1i|1i + |1i|2i + |2i|1i) , (33.10)
3
so that ρAB = |ψihψ|, so S(AB) = 0, but SA , SB > 0. Indeed,
1
ρB = TrA ρ = [2|1ih1| + |2ih2| + |1ih2| + |2ih1|] = ρA = TrB ρ. (33.11)
3
Then
2 2 1 1
SA = SB = ln + ln > 0. (33.12)
3 3 3 3
For an example of S(A + B) = S(A) + S(B), consider noninteracting states, e.g.,
product states. E.g.,
2 1 2 1
ρ = ρA · ρB = |1ih1| + |2ih2| A ⊗ |1ih1| + |2ih2| . (33.13)
3 3 3 3 B
1) Example 1: consider the universal set of gates (NOT, AND, OR, INPUT). Then
to go from (1, 0, 1, 0, 1, 0) to (0, 0, 1, 1, 1, 1) we can consider the two realizations of
the computation:
-(N OT1 , N OT4 , N OT6 ); or
-(1 AN D2 , IN P U T2 , IN P U T3 , 3 OR4 , IN P U T5 , 5OR6 ).
Example 2: Consider the universal set of gates (NOT, AND, OR, XOR). Then
consider the realization of the computation:
-N OT1 ⊗ N OT4 ⊗ 5OR6 .
Therefore, the complexity of the computation is less or equal to 3.
2) For the 5-qubit space, going from
|0i + |1i |0i + |1i |0i + |1i |0i + |1i |0i + |1i
|ψR i = √ , √ , √ , √ , √ (34.1)
2 2 2 2 2
to
|0i − |1i |0i + |1i |0i − |1i |0i + |1i |0i − |1i
|ψT i = √ , √ , √ , √ , √ (34.2)
2 2 2 2 2
can be done without need of an error. Note that |ψT i is not entangled.
We use the complete set of Hadamard gate T , Phase gate P and Toffoli gate T .
Note that
2
2 1 1 1 1 2 1 0
H = √ = 1l , X = 1l , P = ⇒ P 4 = 1l. (34.3)
2 1 −1 0 −1
Then
|0i + |1i |0i − |1i
P2 √ = √ , (34.4)
2 2
so one example of a product of gates giving the computation is
P 2 (1) ⊗ P 2 (3) ⊗ P 2 (5) , (34.5)
so the complexity of the computation is less or equal to 6.
3) A basis OI for a computation acting on a 4-qubit space in the Nielsen approach
is
OI = Oi1 i2 i3 i4 = σi1 ⊗ σi2 ⊗ σi3 ⊗ σi4 , (34.6)
such that the Hamiltonian is
X 3
X
H(s) = Y I (s)OI = Y i1 i2 i3 i4 Oi1 i2 i3 i4 , (34.7)
I i1 ,i2 ,i3 ,i4 =1
120
121 Chapter 34
Then, to show that W (t) spreads away from k, we must calculate it. Since
∞
X (it)n
W (t) = eiHt W e−iHt = [H, ..., [H, W ]...] , (34.10)
n=0
n!
so [H, W ] contains σk−1 , σk , σk+1 . It is then easy to see that [H, [H, W ]] contains
σk−2 , σk−1 , σk , σk+1 , σk+2 , [H, [H, [H, W ]]] contains σk−3 , ..., σk+3 , etc. So indeed,
as t increases, and higher commutator terms become relevant, we have σi further
away from σk becoming relevant, so W (t) spreads.
6) For quantum chaos,
and then F (t) = hψ 0 |ψ̃i is the overlap of the two states, different from hψ̃|ψ̃i = 1 by
a small amount, since initially (at time t → 0) [V, W (t)] → 0 (is very small, leading
to |ψ 0 i ' |ψ̃i), yet it grows with time.
7) From (34.39) in the text, λL ≤ 2πkB T /~, so we infer that λL → 0 as T → 0.
But this doesn’t imply that t∗ → ∞, even though λL ∼ 1/t∗ . The reason is that
(34.37), F (t) ' 1 − eeλL t , is only valid at T 6= 0, yet t∗ is defined from C(t) =
2(1 − ReF (t)) ∼ 1 even at T = 0.
35 Chapter 35
and if |φi = |deadi + |alivei = |ii + |ji, then the probability in this state is
with
X
|...|2 = hi|jihi |j ihψ|jihi|ψi
i,j
123
124 Chapter 35
X
= hi |i ihψ|iihi|ψi =
6 hψ|ψi , (35.7)
i
we first expand the creation and annihilation operators in a (discrete) Fourier series,
L−1
b 2πinj
√n e L ,
X
aj = (35.10)
n=0
L
which preserves the commutation relations: If [ak , a†j ] = δkj , the same is obtained
from [bn , b†m ] = δmn :
L−1 2πi(nk−mj) L−1 2πin(k−j)
e L e L
[ak , a†j ] =
X X
[bn , b†m ] = = δkj . (35.11)
n,m=0
L m=0
L
aa† + a† a (a ± a† )2 α aa† + a† a
α
a2 + (a† )2 ,
H=β ±α =β 1+ ±
2 2 β 2 2β
(35.19)
(so in our case α = αn , β = βn and ± refers to cn,1 , cn,2 ) we can make a Bogoliubov
transformation,
b = α̃a + β̃a† ⇒ b† = α̃∗ a† + b̃∗ a , (35.20)
√ √
For α, β real (like in our case), we define α̃ − β̃ ≡ 1/ ω, α̃ + β̃ ≡ ω, leading to
s s √ √ √ √
1 − α/β 1 + 3α/β ω + 1/ ω ω − 1/ ω
ω1 = , ω2 = ⇒ α̃ = , β̃ = .
1 + 3α/β 1 − α/β 2 2
(35.22)
Then finally we obtain the diagonal Hamiltonian
bb† + b† b (1 + α/β)2 − 2α2 /β 2 bb† + b† b
H= βp ≡ ωn . (35.23)
2 (1 − α/β)(1 + 3α/β) 2
We note that the frequency is the same for both ± signs. We then apply to our
case (for α = αn , β = βn , and ± referring to cn,1 , cn,2 , so b becomes now c̃n,1 , c̃n,2 ),
obtaining the diagonal Hamiltonian
c̃†n,1 c̃n,1 + c̃n,1 c̃†n,1 c̃†n,2 c̃n,2 + c̃n,2 c̃†n,2
L−1
!
X
H= ωn + . (35.24)
n=0
2 2
7) Since thermal particle creation arises from a vacuum state, we can deduce that
energy is not (globally) conserved, so the space is either curved, or topologically
nontrivial. It cannot happen in Minkowski space, which conserves energy.
36 Chapter 36
so the energy is unchanged, basically because the Hermite polynomials obey Hn (−x) =
(−1)n Hn (x), so |ψn (−x)|2 = |ψn (x)|2 , meaning the integrand is odd, and the inte-
gral vanishes. This is in agreement with the exact result.
On the other hand, the wave function changes: we have
X H1,mn
|n(1) i = (0) (0)
|m(0) i. (36.5)
E
m6=n n − Em
127
128 Chapter 36
Z +∞ r r
mωx2 mω mω
= λAn Am dx xe− ~ Hn x Hm x , (36.8)
−∞ ~ ~
and by the same argument, since Hn (−x) = (−1)n Hn (x), if n + m is even, the
integrand is odd, so the integral vanishes. That leaves the case of n + m = 2k + 1.
(0)
Since we also have En = ~ω(n + 1/2), we have
X H1,n+2k+1,n
|n(1) i = |n(0) + 2k + 1i. (36.9)
−~ω(2k + 1)
k
so
Z +∞ n " n+1 #
H1,n+1,n λ 2 2 d 2 2 d 2
=− A2 dy ye−y ey e−y ey e−y .
~ω mω 2 n −∞ dy dy
(36.11)
Then
r
(1) λ mω
hx|n i= (nrs.)hx|n + 1(0) i
mω ~
d
y − dy
r
λ mω
= (nrs.) p ψn(0) (x). (36.12)
mω ~ 2(n + 1)
=λ dr r2 e−µr Rnl
2
(r)
0 Z
∞ 2
2
dr r2 e−2κn r−µr (2κn r)2l L2l+1
= λÑnl n−l−1 (2κn r) . (36.13)
0
But, since
Z ∞
dxe−x(s+1) xµ+β Lµk (x)Lµk (x)
0
A2
1+µ+β
Γ(1 + µ + β)Γ(1 + µ + k) dk 1 F1 2 , 1 + µ+β
2 , 1 + µ; B 2
= (36.14)
,
(k!)2 Γ(1 + µ) dhk (1 − h)1+µ B 1+µ+β
h=0
129 Chapter 36
1+h
where A2 = 4h/(1 − h)2 , B = s + 1−h , we finally obtain
A2
λ(l + 1) dn−l−1 1 F 1 l + 3/2, l + 2, 2l + 2; B 2
En(1) = , (36.15)
n dhn−l−1 (1 − h)2l+2 B 2l+3
h=0
µ
where s = 2κn and
A2 4h 1 + (1 − s)h
2
= 2
, B= . (36.16)
B [1 + s + (1 − s)h] 1−h
For the ground state wave function, we have
(1)
X H1,n0 0
|ψ0 i = (0)
|n0 i ,
(0) 0
(36.17)
0 E
n 6=0 0 − E n 0
(0)
with En0 = E0 /n02 . Due to spherical symmetry, dΩY00 Ylm = δl0 δ0m , so
R
Z
H1,n0 0 = r2 drR10 (r)Rn0 l (r)λe−µr δl0 δm0 , (36.18)
∞
(κ1 /2)3/2
Z
2 −z 1+ µ+κ+1
z dze 2κ 0
n L1n−1 (z) , (36.20)
(κn0 )3/2 0
and since
Z ∞
β −st α Γ(β + 1)Γ(α + n + 1) −β−1 1
dt t e Ln (t) = s 1 F1 −n, β + 1, α + 1; ,
0 n!Γ(α + 1) 2
(36.21)
we obtain
∞ 3
r
(n0 − 1)! n02
(1) λ 3/2
X
0 2κn0
|ψ0 i = (κ1 /2) 2n ×
E0 0 =2
n0 !2n0 n02 − 1 2κn0 + µ + κ1
n
0 2κn0
×1 F1 −n + 1, 3, 2; |n0 00i. (36.22)
2κn0 + µ + κ1
3) To write the second order perturbation theory for E0 , we have
(2)
X |λ(e−µr )0m |2
E0 = (0) (0)
, (36.23)
m6=0 E0 − Em
(0) (0)
where m = (nlm0 ) and Em = E0 /n2 , so
(2)
X |(λe−µr )0;n,l,m0 |2
E0 = (0)
. (36.24)
n,l,m E0 (1 − 1/n2 )
130 Chapter 36
Moreover,
Z ∞
(λe−µr )0;n,l,m = dr r2 λe−µr R10
∗
(r)Rnl (r)δ0l δ0m
0
3/2 Z ∞
1
= r2 dre−κ1 r−µr−κn r L1n−1 (2κn r)
a0 0
3 r
3/2 2κn (n − 1)!
= λ(κ1 /2) 2n ×
2κ n + µ + κ 1 n!2n
2κn
×1 F1 −n + 1; 3; 2; . (36.25)
2κn + µ + κ1
4) For the first excited state, n = 2, we have a degenerate state, with (nlm)
states (200) = |1i, (210) = |2i, (21 + 1) = |3i, (21 − 1) = |4i. Then
Z ∞
H1,αβ = λ(e−µr )αβ = λÑ2l 2
r2 drR2l
2
(r)e−µr δll0 δmm0 = λhe−µr i2l,2l δll0 δmm0 ,
0
(36.26)
so is diagonal in |αi = |1i, |2i, |3i, |4i. We have
(
(1) λhe−µr i20,20 , |αi = |1i
E2 = (36.27)
λhe−µr i21,21 , |αi = rest.
q.e.d.
6) To prove
r
l2 − m2
hlm| cos θ|l − 1 mi = hl − 1 m| cos θ|lmi = , (36.30)
4l2 − 1
we note that it is equal to
Z π Z 2π
∗
= dθ sin θ dφYlm (θ, φ)Yl−1,m (θ, φ) cos θ
Z0 π 0
7) For the Stark effect for n = 3 in the Hydrogenoid atom, we first define the
degenerate basis,
|3; 1i = |3, 0, 0i , |3; 2i = |3, 1, 0i , |3; 3i = |3, 1, 1i
|3; 4i = |3, 1, −1i , |3; 5i = |3, 2, 0i , |3; 6i = |3, 2, 1i
|3; 7i = |3, 2, −1i , |3; 8i = |3, 2, 2i , |3; 9i = |3, 2, −2i. (36.32)
Then we still have
hnl1 m1 |Ĥ1 |nl2 m2 i = dEhl1 m1 | cos θ|l2 m2 ihrinl1 ,nl2 , (36.33)
and still
r
l2 − m2
hlm| cos θ|l − 1, mi = hl − 1, m| cos θ|lmi = , (36.34)
4l2 − 1
and the rest are zero. For hri3l,3l−1 , we define A = hri31,30 and B = hri32,31 , and
in our case we have
r
1
h1| cos θ|2i = h00| cos θ|20i =
3
2
h2| cos θ|5i = h10| cos θ|20i = √
15
1
h3| cos θ|6i = h11| cos θ|21i = √ = h1 − 1| cos θ|2 − 1i = h4| cos θ|7i.(36.35)
15
Then
A
0 √
3
0 0 0 0 0 0 0
√A 0 0 0 2B
√ 0 0 0 0
3 15
√B
0 0 0 0 0 0 0 0
15
0 0 0 0 0 0 √B 0 0
15
Ĥ1 = dE × 0 2B
0 0 0 0 0 0 0 . (36.36)
√
αβ 15
0 0 √B 0 0 0 0 0 0
15
0 0 0 √B 0 0 0 0 0
15
0 0 0 0 0 0 0 0 0
0 0 0 0 0 0 0 0 0
The eigenvalues and eigenstates are found through diagonalization. There are 2
zero eigenvalues, and for the other 7 × 7 matrix, we write the secular equation,
det(H1 − λ 1l) = 0 , (36.37)
which can be solved with Mathematica.
37 Chapter 37
1) For the Hydrogen atom in the ground state, perturbed with λH1,mi = λe−mA+λ̃t ,
for t < 0, the probability to go to the m’th eigenstate is
Pf i (t) = |b(1) 2
n (t)| , (37.1)
where
Z τ
i~b(1)
m (τ ) = dteiωmi t H1,mi (t). (37.2)
0
2
For the Hydrogen atom, ωmi = Em~−E0 = −|E0 | 1/m~ −1 , so
Z τ
−mA+t(λ̃+iωmi ) e−mA (λ̃+iωmi )τ
i~b(1)
m (τ ) = dte = e − 1 ⇒
0 λ̃ + iωmi
e−2mA
2
Pf i (t) = |b(1) 2
m (t)| = 2 2 2 )
e 2λ̃τ
(cos ωτ − 1) 2
+ e 2λ̃τ
sin ωτ (. 37.3)
~ (λ̃ + ωmi
2) For the Hydrogen atom in the ground state, perturbed suddenly (at t = 0)
with Ĥ1 = Kr, to find the transition probability to the n = 2, l = 0 state, we first
find
3
H1,mi = Khri10,20 , Em − Ei = E0 (1/22 − 1) = |E0 | , (37.4)
4
and then substitute in the general formula
4|H1,mi |2 ωmi τ
Pmi (τ ) = sin2 , (37.5)
(Em − Ei )2 2
which at large τ gives
2π
Pmi (τ ) → τ |H1,mi |2 δ(Ef − Ei ) = 0 , (37.6)
~
since Ef 6= Ei .
3) To prove the formula
1 sin2 ax
lim = δ(x) , (37.7)
a→∞ π ax2
2 2
we note that for ax → 0, sin ax sin ax
(ax)2 → 1, so then ax2 → a → ∞. However, for x 6= 0
2 2
(finite) and a → ∞, both sin ax and x are finite, so the limit gives 0. Then, indeed
the limit is proportional to δ(x), it only remains to fix the proportionality constant.
But
Z + Z +∞
sin2 ax sin2 y
dx = dy =π, (37.8)
− ax2 −∞ y2
132
133 Chapter 37
3/2
d ~
r e = 3/2
r 2
dr d cos θe− ~ pr cos θ
(2π~) r (2π~) Z 0 −1 r
∞
1 dr h −r(λ+ ip ip
~ ) − e−r (λ− ~ )
i
= √ e
~ 2π~ 0 − ip ~
2 1
= √ p , (37.14)
~ 2π~ |λ + i ~ |2
so that
2zZ 1
p) = lim √
V (~ ~2
p
. (37.15)
2
λ→0 ~ 2π~ λ +
~2
Ĥ1
Ĥ1 (t) = , t < 0. (37.16)
1 + γ 2 t2
h i
1 1
Writing 1+γ 2 t2 = (1+iγt)(1−iγt) = 21 1−iγt
1 1
+ 1+iγt , we have
τ τ
eiωmi t eiωmi t eoωmi t
Z Z
H1,mi
i~b(1)
m (τ ) = dt 2 2
H1,mi = dt + . (37.17)
−∞ 1+γ t 2 −∞ 1 − iγt 1 + iγt
For τ → ∞, we can add for free a semicircle in the upper-half plane, to close the
134 Chapter 37
contour, since there eiωmi t → 0. For the counterclockwise loop, we get a minus, so
in all, by the residue theorem, we get for the integral
H1,mi iωmi γi
− e . (37.18)
2
Otherwise, for τ < 0, we leave it as it is, and
H1,mi τ
Z
1 1
b(1)
m (τ ) = −i dte iωmi t
+ , (37.19)
2~ −∞ 1 − iγt 1 + iγt
and the transition rate is
τ 2
|H1,mi |2 d
Z
dPf i d (1) 1 1
= |bm (τ )|2 = dteiωmi t + . (37.20)
dτ dτ 4~2 τ −∞ 1 − iγt 1 + iγt
38 Chapter 38
we note that the lifetime is hti = Γ1 = ∆t, and the uncertainty in energy is the
width of the Lorentzian distribution, which is ~Γ, so
1
h∆Ei = ~Γ = ~ ⇒ h∆Eih∆ti = ~. (38.4)
h∆ti
5) The sudden approximation is
(
0, t<0
H1 (t) = , (38.5)
H1 , t≥0
and then
i t 0
Z
ÛI (t, t0 ) = T exp − dt H1,I (t0 ) , (38.6)
~ t0
where
(
0, t<0
H1,I (t) = i
− ~i Ĥ0 (t−t0 )
, (38.7)
e ~ Ĥ0 (t−t0 ) Ĥ1 e , t>0
135
136 Chapter 38
Z t
i X i i
bn (t) = δni − dt1 e ~ En (t1 −t) H1,Snm1 e− ~ Em1 (t1 −t0 )
~ m1t0
Z t1
iX i
= δni − H1S,nm1 e ~ (En −Em1 )(t−t0 )
~ m t0
1
X H1S,nm i
= δni − 1
e ~ (En −Em1 )(t−t0 ) − 1 . (38.9)
m
En − Em1
1
Then
i
X X
|ψi = cn (t)|ψn i = bn (t)e− ~ En t |ψn i , (38.10)
n n
and we obtain
i
X H1S,nm i i
|ψi = e− ~ Ei t |ψi i − 1
e− ~ Em1 t − e− ~ En t |ψn i , (38.11)
m
En − Em1
1
which is just the time-independent result, once we add to it the standard time
dependence.
7) To continue the same to the second order formulae,
Z t Z t
(−i~)2 X
∆b(2)
n (t) = dt1 dt2 H1S,nm1 H1S,m1 m2 ×
2! m ,m t0 t0
1 2
i i
×e ~ (En −Em1 )(t1 −t0 ) e ~ (E
m1 −Em2 )(t2 −t0 )
i
i
X e ~ (En −Em1 )(t−t0 ) − 1 e ~ (Em1 −Em2 )(t− t0 ) − 1
= H1S,nm1 H1S,m1 m2 ,
m ,m
(En − Em1 )(Em1 − Em2 )
1 2
(38.12)
− ~i En t
and, as above, this gets multiplied by e |ψn i. The result matches the time-
independent one,
X 1 X H1,mp H1,pn H 1,nn H1,mn
|n(2) i = − |mi. (38.13)
m
E n − Em E n − Ep E n − E m
p6=n
39 Chapter 39
~2 ~ 2 ~ ~ ~ q ~2
− ∇ + iq A·∇+ A , (39.3)
2m0 m0 2m0
we now add
4 q ~ · p~)~ 2q 2 ~ 2 2 ~ 2 ) + ...
3 (A p2 − (A p~ + (~p · A)
8 m0 c 2 8m30 c2
i q~3 ~ ~ ~ 2 q 2 ~2 ~ 2 ~ 2 ~ · ∇)
~ 2 ) + ...
= ( A · ∇)( ∇ ) − (A ∇ + (A (39.4)
2 m30 c2 4m30 c2
2) To find the expressions for the corrections (due to the above terms) to the
linear and quadratic Zeeman effect, to
q ~ ~ ·B
~ ,
− A · p~ → −µB L (39.5)
m0
giving the linear Zeeman effect, we now add
1 q ~ · p~)~ p~2 ~ · B].
~
+ 3 2
(A p2 → − 2 2 × [−µB L (39.6)
2 m0 c 2m0 c
Also, to
2
q2 ~ 2
m eB
+ A →+ hr2 sin2 θin , (39.7)
2m0 2 2m0
giving the quadratic Zeeman effect, we must add
q2 ~ 2 p~2 + (~ ~ 2 ].
− [A p · A) (39.8)
4m30 c2
The correct treatment will be done later.
3) To show that the splitting of lines according to the Zeeman or the Paschen-
Back rules gives the same number of lines, we remember that the |LSJmJ i basis
137
138 Chapter 39
is equivalent to the |LmL SmS i basis via the Clebsch-Gordan coefficients, so there
must be an equal number of states,
L+S
X
(2J + 1) = (2L + 1)(2S + 1) , (39.9)
J=|L−S|
so the number of lines, for transitions between states, is also the same.
4) For the transition element H1,f i for the Hydrogenoid atom hit by an electro-
magnetic wave, in the dipole approximation, between states with n = 1, l = 0 and
n = 2, l = 1, we first note that for electric dipole radiation, we must have
j 0 = j − 1, j, j + 1 , m0 = m ± 1 , (39.10)
where
E2 − E1 |E1 | 1
ωf i = = 1− , (39.12)
~ ~ 4
and
∗
ρf i = hn = 2, l = 1|ρ̂|n = 1, l = 0i = qψn=2,l=1 (~r)ψn=1,l=0 (~r). (39.13)
Also note that ρ~r = ~j is the electric current. Then we need to calculate (for
~ · ~r = z cos θ, i.e., only z component)
Z
∗
q d3 r~ · ~rψn=2,l=1 (~r)ψn=1,l=0 (~r)
Z ∞ Z
∗ 1
=q r2 dr dΩr cos θR21 (r)Y1m (θ, φ)R10 (r) √
0 4π
= qhlm| cos θ|l − 1m − 1ihri21,10 , (39.14)
q
2 2
and hlm| cos θ|l − 1mi = l4l−m 2 −1 .
~ 10 (r) = R0 (r)∇r,
and ∇R ~ and if ~ · p~ has only a z component, then we obtain z =
10 q r
l2 −m2
cos θ. Also using hlm| cos θ|l − 1mi = 4l2 −1 and rest zero, only for l = 1, m = 0
we get a nonzero result, so we have 12 δl1 . Then finally,
Z ∞
3i 0
M= r2 drj1 (kr)R10 (r). (39.20)
2(2π~)3/2 0
40 Chapter 40
1) To estimate h2|e−a∂x |1i for the harmonic oscillator, using WKB and Bohr-
Sommerfeld quantization, we first remember that the wave function from the B-
S+WKB approximation is
mω 2
x2
ψn ' const. × xn e− 2~ (40.1)
at large x, so
Z +∞
mω 2 mω 2
x2 (x−a)2
h2|e−a∂x |1i = h2|1(x−a)i ' dx x2 e− 2~ (x−a)e− 2~ +intermediate x.
−∞
(40.2)
If a is large, then the above approximation becomes better.
2) For V (x) = − x2 A+a2 and |1i, |2i the first two eigenstates in this potential, then
where
r
x0
2(E − V (x))
Z
dx ~ω1,2
τ≡ , v(x) = , E1,2 ≡ E ± . (40.4)
v(x) m 2
Here x0 is the pole in the upper-half plane in V (x). But there is a unique such
1
pole at x0 = ia, since V (x) = − (x+ia)(x−ia) . Near this pole,
r
A 1
v(x) ' √ , (40.5)
iam x − ia
so
r x0 =ia √
I
iam
τ (x) ∼ dx x − ia. (40.6)
A
But with the parametrization x −√ ia = reiθ , dx = −ireiθ dθ (in the negative θ
direction), the integral is 3 r , and i = 1+i
4 3/2 √ , so
2
r
ma 4 3/2
Imτ ∼ r . (40.7)
2A 3
3) Instantons are not directly associated with the motion of a real particle (clas-
sical, or quantum). The path integral is a sum over quantum paths, and is given by
the instanton approximation, so there is only an implicit relation.
4) Consider x1,2 the minima of U (x), and the series of instanton paths x1 − x2
140
141 Chapter 40
so
∞
X 1
P1n = − 1. (40.9)
n=1
1 − P1
we consider the inverse (Euclidean) potential VE = −α(x2 −a2 ) and classical motion
in it, from x1 = −a to x2 = +a through x0 = 0.
The Euclidean action is
Z [x2 ] " 2 #
1 dx 2 2 2
SE = dtE + α(x − a ) , (40.11)
[x1 ] 2 dtE
which inverts to
√
1 + e2 2αat
x(t) = a √ . (40.14)
1 − e2 2αat
so √
ax 1 + e 2Aat
tan = √ , (40.17)
2 1 − e 2Aat
such that at x → −∞, tan ax π ax
2 = 1, so x = + 2a , whereas for x → +∞, tan 2 = −1,
π
so x = − 2a .
Then the on-shell Euclidean action is (see the previous exercise)
Z +∞ Z +∞ Z +∞ " √ #2
2 2e 2Aat
SE = dt 2V = 2A dt cos ax = 2A dt √ , (40.18)
−∞ −∞ −∞ 1 + e2 2Aat
1 1
1) For the 2-state system with basis |1i, |2i and Hamiltonian H = , the
1 1
first form of the variational method starts with the wave function
X
|ψi = aα |αi = a1 |1i + a2 |2i , (41.1)
α
Z +∞
r
Z +∞ r
~ −y 2 π~
∆11 = dxψ12 (x)
= dye =2
mω mω
Z−∞
+∞
r −∞ Z ∞ r r
~ − 3y2
2
2 π~
∆12 = dxψ1 (x)ψ2 (x) = 2 dye =2
mω 3 mω
Z−∞
+∞
r Z0 ∞ r
~ 2 2 π~
∆12 = dxψ2 (x)ψ2 (x) = 2 dye−2y = √
mω 2 mω
Z−∞
+∞
r Z0 ∞ r r
~ − 5y2
2
2 π~
∆13 = dxψ1 (x)ψ3 (x) = 2 dye =2
mω 5 mω
Z−∞
+∞
r Z0 ∞ r
~ 2 2 π~
∆23 = dxψ2 (x)ψ3 (x) = 2 dye−3y = √
−∞ mω 0 3 mω
Z +∞ r Z ∞ r
~ −4y 2 π~
∆33 = dxψ3 (x)ψ3 (x) = 2 dye =
−∞ mω 0 mω r
Z +∞ r
~ ~ω +∞ d2 ~3 ω √
Z
y2 2
− 2 2 − y2
H11 = dxψ1 (x)Ĥψ1 (x) = dye − 2 +y e = π
−∞ mω 2 −∞ dy r m
Z +∞ r r
~ ~ω +∞ d2 ~3 ω 3π
Z
y2 2
H12 = dxψ1 (x)Ĥψ2 (x) = dye− 2 − 2 + y 2 e−y =
−∞ mω 2 −∞ dy r m r 2
Z +∞ r
~ ~ω +∞ 2
~3 ω 8π
Z
2 d y2
H21 = dxψ2 (x)Ĥψ1 (x) = dye−y − 2 + y 2 e− 2 =
−∞ mω 2 −∞ dy r mr 3
Z +∞ r
~ ~ω +∞ 2
~3 ω π 13
Z
2 d 2
H22 = dxψ2 (x)Ĥψ2 (x) = dye−y − 2 + y 2 e−y =
−∞ mω 2 −∞ dy rm r2 8
Z +∞ r
~ ~ω +∞ 2
~3 ω 5π
Z
y2 d 2
H13 = dxψ1 (x)Ĥψ3 (x) = dye− 2 − 2 + y 2 e−2y =
−∞ mω 2 −∞ dy r m r 2
Z +∞ r
~ ~ω +∞ 2
~3 ω π 1
Z
2 d y2
H31 = dxψ3 (x)Ĥψ1 (x) = dye−2y − 2 + y 2 e− 2 =
−∞ mω 2 −∞ dy r m r 10 5
Z +∞ r
~ ~ω +∞ 2
~3 ω π 11
Z
2 d 2
H23 = dxψ2 (x)Ĥψ3 (x) = dye−y − 2 + y 2 e−2y =
−∞ mω 2 −∞ dy r m r3 4
Z +∞ r
~ ~ω +∞ 2
~3 ω π 7
Z
2 d 2
H32 = dxψ3 (x)Ĥψ2 (x) = dye−2y − 2 + y 2 e−y =
−∞ mω 2 −∞ dy rm 34
Z +∞ r Z +∞ 2
~ ω √ 49
3
~ ~ω 2 d 2
H33 = dxψ3 (x)Ĥψ3 (x) = dye−2y − 2 + y 2 e−2y = π .
−∞ mω 2 −∞ dy m 128
(41.9)
ˆ
q we calculate det(Ĥ − E ∆). In the matrix, we take
For the ground state energy,
3
out the common factor of ~ mωπ , obtaining
q q q q
2E 3 2E 2 5 2E 2
1− ~ωq 2 − ~ω 3 2 − ~ω 5
q
det 2 2E 2 13 2E √1 11
√ − 2E √1
− − = 0. (41.10)
√
3 ~ω 3
q 8 2 ~ω 2 4 3 ~ω 3
2 3/2 1 2E 2 7 2E √1 49 2E 1
5 4 − ~ω 5
√
4 3
− ~ω 3 128 − ~ω 2
145 Chapter 41
2E
Defining a ≡ ~ω ,
the equation becomes, explicitly,
r r !r
13 a 49 a 5 2 2 7 a
(1 − a) √ −√ − + −a (a − 1) √ −√
8 3 2 128 2 2 5 3 4 3 3
r r ! r r r ! r
3 2 11 a 2 1 5 2 13 a 2 1
+ −a √ −√ −a − −a √ −√ −a
2 3 4 3 3 5 10 2 5 8 2 2 3 10
r r !r
3 2 2 49 a 7 a 11 a
− −a (1 − a) − − (1 − a) √ −√ √ −√ = 0. (41.11)
2 3 3 128 2 4 3 3 4 3 3
We note that, among the 6 terms, 4 are canceled by putting a = 1, which is the
exact ground state of the harmonic oscillator (E = ~ω
2 ), and we are left with (for
a = 1)
r " r r ! r r ! #
2 9 5 2 5 3 2 7
− √ − − √ , (41.12)
5 10 2 5 8 2 2 3 4 3
which means that the exact ground state is very close to the lowest solution of the
approximate equation.
2
3) For the practical variational method, with trial wave function ψa (x) = e−ay ,
in order to find the ground state energy, we write
~ω ∞ 2
d2 2
dye−ay − dy 2
e−ay
R +∞ R
2 + y
∗
−∞
dxψ (x)Hψ(x) 2 0
hEiψ = R +∞ = R∞
−∞
dxψ ∗ (x)ψ(x) 0
dye−2ay2
" √ #
~ω 2
= (1 − 4a2 ) + 2a . (41.13)
2 4a
so that
~ω 7a 5
hEiψ = + 2 . (41.17)
2 3 8a
Minimizing over a, we obtain
1/3
∂hEiψ 30
= 0 ⇒ 2a = , (41.18)
∂a 7
and substituting back into the energy, we find
1/3
7 30
hEiψ = ~ω , (41.19)
8 7
which is not too close to the correct value 52 ~ω.
2
5) For the potential V = kx2 + α|x|3 , and the trial wave function ψ(x; a, b) =
2
|y|a e−by , we have the Hamiltonian
r !
~ω d2 2 3 2α ~
H= − 2 + y + |y| , (41.20)
2 dy k mω
q
and defining A ≡ 2α k
~
mω , we find
R +∞
a −by 2 d2 3
~ω −∞
dy|y| e − dy 2 + y 2
+ A|y| 3
|y|a e−by
hEiψ = R +∞
2 dy|y|2a e−2by2
−∞
1 − 4b2
~ω a(a − 1)2b
= − + 2b(2a + 1) + (a + 3/2)(a + 1/2)
2 a − 1/2 2b
A
+ (a + 5/2)(a + 3/2)(a + 1/2) . (41.21)
(2b)3/2
The equations for the minimum of the energy (to find the correct approximation)
are
∂hEiψ ∂hEiψ
= = 0. (41.22)
∂a ∂b
6) For the details for the minimization of the trial wave function R(r) = Ae−r/a
for the Hydrogen atom, we write
~2 ~ 2 Ze20 ~2
2
~2 l(l + 1) Ze20
d 2 d
Ĥ = − ∇ − =− 2
+ + − , (41.23)
2m r 2m dr r dr 2mr2 r
so that
Ze20
h 2 2 i
R∞ −r/a ~2 l(l+1)
0
r 2
dre − ~
2m dr 2
d
+ 2 d
r ∂r + 2mr 2 − r e−r/a
hEiψ = R∞
0
r2 dre−r/a
Z ∞ 2
2~2 2Ze20 1 2~2 l(l + 1) 1
1 ~
= t2 dte−t − 2
+ − +
2 0 2ma ma2 a t ma2 t
2 2
~ 1 Ze0
= + + 2l(l + 1) − . (41.24)
ma2 2 a
147 Chapter 41
Ze20
R∞ h 2 2 i
r2 d 2 d 1
0
dr 2
r +b 2 − ~
2m dr 2 + r dr − r r 2 +b2
hEiψ = R∞ 1
0
r2 dr (r2 +b 2 )2
2 Ze2
− ~ 2 (2I1 − 10I2 + 8I3 ) − 2b0
= 2mb π , (41.30)
2 − I1
R∞ dx
where we have defined In = 0 (x2 +1) n+1 . Then, when minimizing, we get
∂hEiψ 2~2
=0⇒b=− (2I1 − 10I2 + 8I3 ) , (41.31)
∂b mZe20
and when substituting back into the energy, we find the minimum (approximation
for the ground state energy)
Ze20
hEi0 = − . (41.32)
4b π2 − I1
42 Chapter 42
L32 (z) = 60(z 2 − 10z + 20) , L13 (z) = −4(z 3 − 12z 2 + 36z − 24). (42.2)
where we have shown explicitly what happens only for the non completely filled
orbitals.
3) For the element O (Oxygen), with 1s2 , 2s2 , 2p4 (see above), considering that
s has l = 0 and p has l = 1, and in the 2p4 , two electrons are parallel, and two are
antiparallel, so S ~ + 1/2
~ = 1/2 ~ + ~0 = ~1 (the filled s orbitals don’t contribute) and
~ = ~1 + ~1 + ~1 + ~1 (the filled orbitals don’t contribute), but two of the ~1’s cancel, so
L
~
L = 2, and then J = 1, 2, 3 for the LS coupling.
On the other hand, for the j − j coupling, again the filled s orbitals don’t con-
tribute, and we have ~j1 = ~1 + 1/2 ~ = 3/2,~ ~j2 = ~1 − 1/2
~ = 1/2,~ ~j3 = 1/2
~ + ~0 = 1/2,
~
~ + ~0, and then ~j1 + ~j2 = ~1 or ~2, and ~j3 + ~j4 = ~1, and when finally summing
~j4 = 1/2
the two to get J, ~ again we have J = 1, 2, 3 (note that J~ = ~0 is excluded, since there
is total spin).
4) The method of atomic orbitals for O, in the basis of only filled orbitals gives
3
X
|ψa i = Ck |φk i , (42.4)
k=1
where
|φ1 i = |1s2 i , |φ2 i = |2s2 i , |φ3 i = |2px i. (42.5)
148
149 Chapter 42
Then the minimum energy is found, as usual, from the secular equation
3
X
Ck (Hik − Eδik ) = 0 ⇒ det(Hik − Eδik ) = 0. (42.6)
k=1
5) For O2 , where each Oxygen has the orbital structure (1s2 , 2s2 , 2p4 ), the
method of molecular orbitals starts with molecular orbitals (the notation means
the electron is in the first orbital with respect to the first nucleus, and the second
orbital with respect to the second)
φ113366 11 12 = (φ1 )2 (φ3 )2 (φ6 )2 φ11 φ12 , φ122366 11 12 = φ1 (φ2 )2 φ3 (φ6 )2 φ11 φ12 , etc.
(42.8)
6) The LCAO method for O2 applied to each atomic orbital in O is:
The molecular orbital is
~ 1 ) + Ci2 χ2 (~ri , R
φai i (~ri ) = Cir χr (~ri , R ~ 2) , (42.9)
for instance
1) For a nucleus spinning around an axis, it will deform to be an ellipsoid, but in the
liquid droplet model we would still have the volume proportional to A, V = V0 A (V0
being the volume of a single nucleon). This will replace the particular R = r0 A1/3 ,
valid only for a sphere.
2) The rotating nucleus become an ellipsoid. To find the excentricity, we start
from the observation that the difference in pressure between the exterior and the
interior of the curved surface of the ellipsoid is
1 1
∆p(out − in) = σ + , (43.1)
R1 R2
where σ is the surface tension and R1,2 are the two local radii of curvature of the
two-dimensional surface. With respect to the axis of rotation, we define a North
Pole A, where the distance to the center O is smallest, and an equator, with a point
B on it, where the distance to the center O is largest.
Then at the North Pole, the two local radii of curvature are equal, R1 = R2 =
R + ∆R, whereas at the Equator, the radius in the Equator direction is still R1 =
R + ∆R, but the radius in the meridian direction (towards the North Pole) is
R2 = R − ∆R. That means that there is an extra pressure at the Equator, and that
is due to the effect of the rotation, pushing matter outwards. But the pressure at
the center O is the same, whether we integrate it from A to O, or from B to O.
Then, when comparing the two ways to integrate, we should get the same result
at the center. That means that the integral of the centrifugal pressure from B to
O should equal the extra pressure ∆p. We have for the centrifugal pressure
d2 F
d2 F = d2 mω 2 r = ρ · 2πrdrdy · ω 2 r ⇒ dp = = ρω 2 rdr. (43.2)
2πrdy
Integrating, we have
R2
∆Pextra = ρω 2 . (43.3)
2
But then,
R2
2σ 1 1
∆p(A − O) = = ∆p(B − O) = σ + − ρω 2
R + ∆R R + ∆R R − ∆R 2
2
2σR 2R
= 2 − ρω . (43.4)
R − ∆R2 2
That gives
∆R R2 ∆R ρR3 ω 2 3ω 2
2σ 2
' ρω 2 ⇒ = = M. (43.5)
R 2 R 4σ 16πσ
151
152 Chapter 43
4) For a spherical square well with V = 0 for r < R, the solution for r < R is
r
χnl (r) 2m(E + V0 )
= Nn,l jl (kr) , k = , (43.19)
r ~2
and the solution for r ≥ R is
r
χnl (r) 2mE
= Ñn,l jl (k̃r) , k̃ = . (43.20)
r ~2
Then the matching equations at r = R are
r2
r
VN (r) → −V0 1 − 2 1− , (43.24)
R1 R2
whereas for r → ∞ (r R2 ), we have VN (r) → 0, so we have a better approxima-
tion to the real case. That means that at large n, we are closer to the reality, with
respect to the harmonic oscillator approximation considered in the text. Instead of
being linear, En = ~ω(n + 3/2) ∼ n, they will accumulate, En − En+1 → 0.
6) We have
1 dVN (r)
Hint,spin−orbit = −const. (2~s · ~l) , (43.25)
r dr
and
mω12
V0 1 dVN (r) 2V0 r
= , = 2 e−r/R2 1− , (43.26)
R12 2 r dr R1 2R2
154 Chapter 43
while, as before,
(
l, j = l + 1/2
2~s · ~l = . (43.27)
−l − 1 , j = l − 1/2
Then,
r l
Hint,spin−orbit = −(const.0 ) × e−r/R2 1− × , (43.28)
2R2 −l − 1
so this is still < 0 for j = l + 1/2 and > 0 for j = l − 1/2, but now only for
r < R2 , that is, for the first numbers n. Otherwise (for larger numbers n), the two
are switched, and much smaller.
7) For the pairing energy for a 2-particle delta function potential, for 2 nucleons
in the ground state 1s1/2 , l = 0, s = 1/2, so j = 1/2, and from equations (49.47)
and (43.48) in the text, we obtain
Z
2g dr
∆E = h1/2, 1/2, 0, 0|V̂ |1/2, 1/2, 0, 0i = |φ10 (r)|2 |φ10 (r)|2 . (43.29)
8π r2
But
√ 1/2 √
α 2α 1 1/2 2 2 4α 2α − α2 r2
φ10 (r) = 1 √ 31 √ L0 (α r ) = √ e 2 , (43.30)
2 π 22 π π 3
p
where α = mω/~, so
gα7 √ √
∆E = 5
256 2(− π). (43.31)
9π
44 Chapter 44
and we match against the right-hand side of the first Bloch equation, which is
1 ω0 − ω
− V0 2i sin ωt cos ωt Q(0) + P (0) sin Ωt
i~ Ω
ω0 − ω 2V0 ω0 − ω
+ N (0) − Q(0) (cos Ωt − 1)
Ω ~Ω Ω
2V0 ω0 − ω
+ sin ωt P (0) cos Ωt + N (0) − Q(0) sin Ωt
~Ω Ω
+2iV0 cos ωt {cos ωtP 0 (t) − sin ωtQ0 (t)}]
2V0 0
=− P (t)
~
2V0 2V0 ω0 − ω
=− P (0) cos Ωt + N (0) − Q(0) sin Ωt , (44.3)
~ ~Ω Ω
where on the second line we noticed a Q0 (t) and on the third a P 0 (t), which then
add up to the simple form on the fifth line.
2) Continuing with the second and third Bloch equations, we obtain first for the
left-hand side of the second Bloch equation
dQ(t)
= sin ωt(Ṗ 0 − ωQ0 ) + cos ωt(Q̇0 + ωP 0 )
dt
2V0 ω0 − ω
= sin ωt −Ω sin ΩtP (0) + Ω cos Ωt N (0) − Q(0)
~Ω Ω
ω(ω0 − ω)
−ωQ(0) − P (0) sin Ωt
Ω
ω(ω0 − ω) 2V0 ω0 − ω
+ N (0) − Q(0) (cos Ωt − 1)
Ω ~Ω Ω
2V0 ω0 − ω
+ cos ωt (ω0 − ω)P (0) cos Ωt + (ω0 − ω) sin Ωt N (0) − Q(0)
~Ω
Ω
2V0 ω0 − ω
+ωP (0) cos Ωt + ω sin Ωt N (0) − Q(0) , (44.4)
~Ω Ω
155
156 Chapter 44
the equivalent of Fermi’s golden rule is derived as follows. From eq. (44.24) in the
text,
1 t 0
Z
(iω−γ)t0 † (−iω−γ)t0 0
c(1)
n (t) = dt V 0,ni e + V 0,ni e eiωni t
i~ 0
1 1 − e[i(ω+ωni )−γ]t 1 − e[i(ωni −ω)−γ]t †
= V0,ni + V0,ni . (44.9)
~ ω + ωni + iγ −ω + ωni + iγ
That means that ~ωni = En − Ei is now replaced with
~ωni ± ~ω + iγ = En − Ei ± ~ω + iγ , (44.10)
leading to a modified Fermi’s golden rule of
!Z
2π |V0,ni |2
Z
dPni X
= † 2 dE ... ρ(En , ...)δ(En − Ei ± ~ω + iγ). (44.11)
dt ~ |V0,ni |
4) We want to find a limit for which the first order perturbation theory in the
harmonic potential is consistent with the exact, two-level calculation. It is easy to
see that this is V0 → 0 (but keep small perturbations, so that the approximation is
exact), and ω → ωni = E2 −E~
1
.
Then, for first order perturbation theory,
1 1 − ei(ω+ωni )t 1 − ei(ωni −ω)t †
(1)
cn (t) = V0,ni + V0,ni
~ ω + ωni −ω + ωni
(ωni −ω)t (ωni −ω)t (ωni −ω)t
ei 2 e−i 2 − ei 2
†
→ −V0,ni , (44.12)
~ω − (En − Ei )
where we have neglected the first term in the limit, and so the transition probability
becomes
† 2
2iV0,ni sin(ω − ωni )t/2
Pn(1) (t) = |c(1) 2
n (t)| →
~ω − (En − Ei )
(2V0,ni )2 sin2 ω − En −E t
i
~ 2
= . (44.13)
|~ω − (En − Ei )|2
This gives the same result as from the exact two-level system in the same limit
V0 → 0, ω → ωni , since then Ω → ω − ω0 , and
2
(2V0 )2
2V0 Ωt E2 − E1 t
P2 (t) = sin2 → sin2
ω − . (44.14)
~Ω 2 (~ω − (E2 − E1 ))2 ~ 2
5) For a possible quantum term αA ~ 2 , we want to analyze in terms of photons,
like in the text for the linear term in A. ~ We have
h
~ ~0 0 ~ ~0 0
~j (~k) · ~l (~k 0 ) aj (~k)al (~k 0 )ei(k+k )~r−i(ω+ω )t + a†j (~k)a†l (~k 0 )e−i(k+k )~r+i(ω+ω )t
X
αA ~2 ∝
~
k,~
k0 ,j,l
158 Chapter 44
i
~ ~0 0 ~ ~0 0
+aj (~k)a†l (~k 0 )ei(k−k )~r−i(ω−ω )t + a†j (~k)al (~k 0 )e−i(k−k )~r+i(ω−ω )t . (44.15)
√ √
Now, since as before we have hN − 1|a|N i = N and hN + 1|a† |N i = N + 1,
we have two terms,
p
hN − 2|aj al |N i = N (N − 1) , (44.16)
corresponding to an absorption of two photons (at the same time), and
hN + 2|a†j a†l |N i = (N + 1)(N + 2)
p
(44.17)
corresponding to an emission of two photons (at the same time), but we also have
the terms
hN |a†j al |N i = N , and hN |aj a†l |N i = N + δjl δ(~k − ~k 0 ) , (44.18)
where nothing happens.
6) If there are no photons, Nj (~k) = 0 in the radiation field, we interpret the rate
of spontaneous emission as describing emission of the photon from the vacuum. The
limitations of this situations are that this is a highly quantum process, so it has no
classical limit, and therefore must be treated fully within Quantum Field Theory.
7) In deriving the Planck formula, we used the Boltzmann distribution for atoms,
which seems counterintuitive, since we describe a quantum process, but we consid-
ered many atoms at equilibrium (the distribution is over atomic states), in which
case we have a classical distribution for them.
45 Chapter 45
2
1) For the potential V (x) = V0 e−αx , for E ≤ V0 , we want to calculate the first
two terms in the perturbative expansion for ψ1 (x) (∼ e−kx at x → −∞) and ψ2 (x)
(∼ eikx at x → +∞). We have the self-consistency equation
1 x
Z
ψ1 (x) = eikx + dx0 sin[k(x − x0 )]U (x0 )ψ1 (x0 ) , (45.1)
k −∞
(0)
where U (x) = 2mV (x)/~2 , solved perturbatively as ψ1 = eikx and
Z x
(1) 2mV0 1 0
ik(x−x0 ) −ik(x−x0 )
02 0
ψ1 (x) = 2
dx e − e e−αx eikx
~ 2ik −∞ Z
x Z x
2mV0 1 ikx 0 −αx02 −ikx 0 2ikx0 −αx02
= e dx e − e dx e . (45.2)
~2 2ik −∞ −∞
Similarly,
Z +∞
1 0 0
ψ2 (x) = eikx − dx0 eik(x−x ) − e−ik(x−x ) U (x0 )ψ2 (x0 ) , (45.3)
2ik x
(0)
so ψ2 (x) = eikx and
1 2mV0 +∞ ik(x−x0 )
Z
(1) 0 02 0
ψ2 (x) = − 2
e − e−ik(x−x ) e−αx eikx
2ik ~ x
1 2mV0 ikx +∞ 0 −αx02
Z Z +∞
−ikx −αx02 ikx0
=− e dx e −e e e .(45.4)
2ik ~2 x x
2) We have
∂x2 G2 (x, x0 ) = −k sin[k(x − x0 )]θ(x0 − x) + cos[k(x − x0 )]θ0 (x0 − x)
= −k 2 G2 (x − x0 ) + δ(x − x0 ) , (45.5)
so G2 (x, x0 ) is also a Green’s function for the Schrödinger operator, and moreover
∂x2 (G1 − G2 ) = −k 2 (G1 − G2 ) , (45.6)
so G1 − G2 is a solution to the homogenous Schrödinger operator.
3) For the Lippman-Schwinger equation in the momentum representation, we
multiply by hk 0 | the abstract equation
1
|ψi = |ki + V̂ |ψi , (45.7)
E + i − Ĥ0
obtaining
1
ψ(k 0 ) = δ(k − k 0 ) + hk 0 | |k 00 ihk 00 |V̂ |k̃ihk̃|ψi
E + i − Ĥ0
159
160 Chapter 45
Z
0 1
= δ(k − k ) + dk̃V (k 0 , k̃)ψ(k̃). (45.8)
E + i − Ĥ0 (k 0 )
The perturbative solution gives
1 1 1 1
ψ(k 0 ) = δ(k − k 0 ) + hk 0 | + + V̂ + ... V̂ |ki
E + i − ĤZ
0 E + i − Ĥ0 E + i − Ĥ0 E + i
0 1 0
= δ(k − k ) + 0
dk̃V (k , k̃)δ(k̃ − k)
Z Z E + i − H0 (k )
˜ 0 1 ˜ ˜
+ dk̃ dk̃V (k , k̃) V (k̃, k̃)δ(k̃ − k) + ... (45.9)
E + i − H0 (k̃)
4) For V = V0 , |x| ≤ L/2, V = 0, |x| > L/2, and E ≤ V0 , we have:
~2 d2
-For |x| > L/2, the Schrödinger equation is − 2m dx2 ψ(x) = Eψ(x), so the solu-
tion in this region is
~2 k 2
ψ(x) = Aeikx + Be−ikx , = E. (45.10)
2m
2 2
~ d
-For |x| ≤ L/2, the Schrödinger equation is − 2m dx2 ψ(x) = (E − V0 )ψ(x), so the
solution in this region is
~2 κ2
ψ(x) = Ceκx + De−κx , = |V0 − E|. (45.11)
2m
For ψ1 (x): For x < −L/2 (incoming), we have ψ1 (x) = Aeikx , so from the
matching conditions at x = −L/2, we get
Then, at x → +∞,
F
ψ1 (x) = E eikx + e−ikx . (45.16)
E
For ψ2 (x): ψ2 (x) = Eeikx at x → +∞.
Matching at x = +L/2, we have
(ik−κ)L κ + ik
1 + e2κL κ−ik
κ+ik
A = Ee . (45.20)
2κ 1+ B A e ikL
At x → −∞, we have
B
ψ2 (x) = A eikx + e−ikx . (45.21)
A
Then the transfer matrix is
α(k) β ∗ (k)
K= , (45.22)
β(k) α∗ (k)
where
ˆ ~ · ~σ
α
= ei∆ = eiα0 cos |~
α| 1l + i sin |~
α| , (45.24)
|~
α|
and by identifying the two forms, we get
1 iα0 α2 iα0 β∗ + β
= e cos |~
α | , α 3 = 0 , e sin |~
α | =
α∗ |~
α| 2α∗
α1 iα0 β∗ − β
e sin |~ α| = −i , (45.25)
|~
α| 2α∗
so further, α3 = 0, and defining β = β1 + iβ2 ,
except in order for that to be true, we note that (Pj,j+1 − 1)|x1 , x1 + 1i, for j = x1
gives 0 instead of the needed
so we need that
ei(p1 x1 +p2 (x1 +1)) + S(p2 , p1 )ei(p1 (x1 +1)+p2 x1 )
×(|x1 + 1, x 1 + 1i + |x1 , x1 i − 2|x1 , x1 + 1i)
= eix1 (p1 +p2 ) eip2 + S(p2 , p1 )eip1 = 0 , (45.31)
which is satisfied by
1
2 cot p22 − 1
2 cot p21 + i
S(p2 , p1 ) = . (45.32)
1
2 cot p22 − 1
2 cot p21 − i
1) For V = V0 > 0 for r ≤ R and V = 0 for r ≥ R, and E > V0 , l > 0, and a square
integrable solution, we have the Schrödinger equation
∂ 2 ψ 2 ∂ψ l(l + 1)
+ − ψ + k2 ψ = 0 , (46.1)
∂r2 r ∂r r2
where weq
have implicitly assumed ψq= REl (r)Ylm (θ, φ), and REl = χEl /r, ρ = 2kr,
and k = 2mE
~2 for r > R and k̃ = 2m(E−V ~2
0)
for r < R. Then we have
∂ 2 χ(ρ)
l(l + 1) 1
+ − + χ(ρ) = 0 , (46.2)
dρ2 ρ2 4
which has the solution
χ(ρ)
REl (ρ) = = jl (ρ/2) = jl (kr) , (46.3)
r
with the correct, square integrable behaviour at r =R0, since jl (kr) ∼ rl for r → 0,
and square integrability of ψ means 0 |χ| drdΩ = 0 |ψ|2 r2 drdΩ < ∞, so ψ goes
2
R
and
1 ±i(kr− π (l+1))
jl (kr) ± iyl (kr)(−1)l+1 ∼ e 2 . (46.8)
kr
The matching conditions for the wave function at r = R are
out of which we find α, β, and then, using the asymptotics at infinity of the spherical
Bessel functions, we have
αjl (kr) + βyl (kr) = A(jl (kr) + i(−1)l yl (kr)) + B(jl (kr) − i(−1)l yl (kr)) ⇒
α − βi(−1)l+1 α + βi(−1)l+1
A= , B= . (46.10)
2 2
2) The relation (46.15),
~ 2π eikr 2π e−ikr
eik·~r ' δ(~nk − ~nr ) − δ(~nk + ~nr ) , (46.11)
k r k r
is understood, as shows its proof, as a distribution in ~nr , i.e., by d2~nr f (~nr )× it,
R
and its says that the bulk of the integral comes from the ~k parallel or antiparallel
to the ~r region.
3) If f~k (~nr ) in
~ eikr
u~+ (~r) ' eik·~r + f~k (~nr ) (46.12)
k r
would be independent of ~nr , it would not contradict unitarity, since then we would
still have
ik ik
Ŝ = 1l + f~k ⇒ Ŝ † = Ŝ −1 = 1l − f~ , (46.13)
2π 2π k
and Ŝ would still be unitary. In other words, the scattered wave is not larger than
the incoming wave. Note that the optical theorem is
Z
k
Imf~k = |f~k |2 d2~nr , (46.14)
2π
and it is still OK.
R
4) For the case at exercise 1), since we have by definition A = S · B, and more
precisely
ik
Ŝ = 1l + f, (46.15)
2π
it follows that, if f is small (so we can ignore f 2 terms), we have
and
Z Z
σtot = d2~nr |f |2 = d2 Ω|f |2 . (46.17)
166 Chapter 46
5) Considering
e−ikr
ikr
2π e 2π
ψ= δ(~nk + ~nr ) + a + δ(~nk − ~nr ) , (46.18)
k r k r
with a, b ∈ R, we can interpret it as a scattering solution. Then
ik ik
f = A − B = a − b , Ŝ = 1l +f = 1l + (a − b) , (46.19)
2π 2π
R
and A = S · B works, except for the b(a − b) term, which can be neglected. Then
Z Z
σtot = d ~nr |f | = d2 Ω|f |2 = 4π|a − b|2 .
2 2
(46.20)
The case with a, b ∈ C doesn’t work, since then we have no optical theorem.
6) To calculate the generalized Green’s function G0 (z; ~r, ~r0 ) for complex energy
z ∈ C, we write
1 1
Ĝ0 (z) = ⇒ Ĝ0 (z; ~r, ~r0 ) = h~r| |~r0 i. (46.21)
z − Ĥ0 z − Ĥ0
k2 2 ~2 k 2
We then define, analogously to E = ~2m , z = 2mz , then follow everything up to
eq. (16.61) in the text, with k 2 ± i replaced by kz2 ∈ C, so
Z +∞ 0 0
0 1 eiq|~r−~r | − e−iq|~r−~r |
G0 (z; ~r, ~r ) = − 2 qdq . (46.22)
8π i|~r − ~r0 | −∞ q 2 − kz2
This has poles at q = ±kz ∈ C. If Im kz > 0, then this is like G+
0 , if Im kz < 0,
this is like G−
0 , so finally
1 0
G0 (z; ~r, ~r0 ) = − 0
e±ikz |~r−~r | , (46.23)
4π|~r − ~r |
with the ± corresponding to Im kz > or < 0, so that always the exponential is
decreasing.
7) For the equivalent of eq. (46.69) in the text, i.e.,
e±ikr 2m
Z
~˙ ~0 0
ψ± (~r) = eik~r − d3~r0 e∓ik ·~r V (~r0 )ψ± (~r0 ) (46.24)
r 4π~2
for G0 (z; ~r, ~r0 ), we substitute it in
Z
2m
ψz (~r) = φz (~r) + 2 d3~r0 G0 (z; ~r, ~r0 )V (~r0 )ψz (~r0 ) , (46.25)
~
~
and use φz (~r) = eikz ~n·~r ≡ eikz ·~r , so
Z ±ikz |~ r0 |
r −~
~
kz ·~ 2m 3 0e
ψz (~r) = e r
− d ~r V (~r0 )ψz (~r0 ). (46.26)
4π~2 4π|~r − ~r0 |
0 0
Using e±ikz |~r−~r | ' e±ikz r e∓ikz ~nr ·~r , we finally find
2m e±ikz r
Z
i~
kz ·~ ~0 0
ψz (~r) = e r
− 2
d3~r0 e∓ikz ·~r V (~r0 )ψz (~r0 ) , (46.27)
4π~ r
where ~kz0 = kz ~nr .
47 Chapter 47
1) For V (r) = −V0 δ 3 (~r), we want to calculate the first 2 terms in the Born series.
We note that the potential is spherically symmetric so, with q = |~k 0 − ~k|, we have
Z
2m ~ 0 ~ 2m 0 ~0 ~
(1)
f =− 2
V (k − k) = − 2
d3~r0 ei~r ·(k −k) [−V0 ]δ 3 (~r)
4π~ 4π~
2m
= V0 , (47.1)
4π~2
so in first order we have
so it is divergent.
2) For the potential V (r) = A/r2 , we want to calculate dσ/dΩ in the Born
approximation. We have
2m
f (1) = − V (~k 0 − ~k) , (47.4)
4π~2
Then we have
2 2 2
dσ (1) 2π 2 A
2m 2m
= |f (1) | = [V (q)]2 = . (47.6)
dΩ 4π~2 4π~2 q
3) To understand how is it possible for the Born approximation to quantum
mechanics for scattering in Coulomb potential to give the same result as classical
scattering, we note that what this means is that the classical result is exact, and
there are no quantum corrections to it. This means the Coulomb case is a very
special case, and it is not expected to be true in general.
4) For the potential V (r) = A/(r2 + a2 ), we want to calculate the first two terms
in the Born series. By spherical symmetry, we have
2m
f (1) = − V (q) , (47.7)
4π~2
and
4π ∞ 0 0 A
Z Z
0 A
V (q) = d3~r0 ei~r ·~q 2 = r dr 02 sin qr0
r + a2 q 0 r + a2
4πA +∞ zdzeiz
Z
=
2iq −∞ (z + iqa)(z − iqa)
2π 2 A iqa
= e , (47.8)
q
where again we have extended the integral over R to an integral over a closed
contour, by adding for free the semicircle at infinity in the upper-half plane, and
then evaluated it using the residue theorem as 2πi times the residue in the upper-
half plane, specifically at z = +iqa. We note that the result is consistent with the
result at exercise 2, since the a → 0 limit gives V = A/r2 , and then the V (q) ’s
match.
Then
2m 2π 2 A iqa
f (1) = − e , (47.9)
4π~2 q
and the first term in the Born series is
(1) e±ikr (1)
ψ± = f , (47.10)
r
whereas the second term in the Born series is
!
Z ±ik|~ k0 |
r −~
(2) 2m e (1)
ψ± = d3~r0 − ψ± (~r0 )V (~r0 )
4π~2 4π|~r − ~r0 |
2 2 iqa Z 0 0
e±ik|~r−~r | e±ikr
2m A e 3 0
= d ~
r , (47.11)
4π~2 q |~r − ~r0 |(r02 + a2 ) r0
0 0 0
and, using e±ik|~r−]~r | ' e±ikr e∓i~r ·~r as in the text, we obtain
2 2 iqa ±ikr Z ~0 0 0
e∓ik ·~r ±ikr
(2) 2m A a e 3 0
ψ± ' d ~r 0 0
4π~2 q r r (r + ia)(r0 − ia)
169 Chapter 47
2 2 iqa ±ikr 0
2π ∞ 0 1 − e±2ikr
Z
2m A e e
= dr
4π~2 q r k 2i(r0 + ia)(r0 − ia)
2 2 iqa ±ikr 0 Z ∞
2m A e e 1 1 dz dz
= 2π −
4π~2 q r 2i 2ika 0 z − ika z + ika
1 ∞ dze±2iz
Z
−
2i 0 (z + ika)(z − ika)
2 2 iqa ±ikr
2m A e e πi 1 −2ka
= 2π + e , (47.12)
4π~2 q r 2ka 2
where in doing the second integral over dz we again added a semicircle at infinity
and used the residue theorem.
−µr
5) For V (r) = V0 e r , the Lippman-Schwinger equation for the T-matrix is
T̂ = V̂ + V̂ Ĝ0 T̂ , (47.13)
and in coordinate space, acting on momentum space, it is
h~r|T̂ |~ki = h~r|V̂ |~ki + h~r|V̂ Ĝ ~
Z 0 T̂ |ki −µr
e−µr i~k·~r Vo e
= V0 e + d3~r0 G0 (~r, ~r0 )h~r0 |T̂ |~ki , (47.14)
r r
and Z
h~r|T̂ |~ki = d3~k 0 ei~k · ~rT (~k, ~k 0 ). (47.15)
6) For V (~r) = −V0 δ 2 (~r), as we saw V (~k) = −V0 . The Lippman-Schwinger equa-
tion for T̂ is
( 1l − V̂ Ĝ0 )T̂ = V̂ ⇒ T̂ = ( 1l − V̂ Ĝ0 )−1 V̂ , (47.16)
so
1 1 1
h~k|T̂ |~k 0 i = h~k| V̂ |~k 0 i = V (~k 0 )h~k| |~k 0 i = −V0.
1 − V̂ Ĝ0 1 − V̂ Ĝ0 δ(k − k ) + V0 G0 (~k, ~k 0 )
~ ~ 0
(47.17)
7) There is no domain of validity of the Born approximation in the case of the
Coulomb potential, since the Born approximation is always valid in the case of the
Coulomb potential.
48 Chapter 48
2mV0 R2 X
1 1
f (1) (~k, ~k 0 ) = − (−1)m
(qR)2m
+
~2 (qR) (2m)! (2m + 1)!
m≥0
r 2m
2mV0 R2 (1 − cos θ)m
2 X kR
=− 2 (−1)m √ (2m + 2).
~ 2kR 1 − cos θ 2 (2m + 1)!
m≥0
(48.7)
It is complicated in practice to match this against the general expansion, so we
will not continue.
3) We have
Sl (k) = e2iδl (k) (48.8)
170
171 Chapter 48
and
2
dσ 2mV0
= |f (1) (~k, ~k 0 )|2 = (qR cos qR + sin qR)2 . (48.9)
dΩ ~2 q 2
4) If δl (k) ∈ R, we want to deduce something about Tl (k):
e2iδl (k) = Sl (k) = 1 − 2πiTl (k) ⇒ Sl† (k) = e−2iδl (k) = (Sl (k))−1 , (48.10)
so unitarity means
Sl Sl† = 1l ⇒ (1 + 2πiTl† )(1 + 2πiTl ) = 1 ⇒ 2πi(Tl† (k) − Tl (k)) + (2π)2 |Tl (k)|2 = 0.
(48.11)
5) The scattering length is
2mV0
a = − lim a0 (k) = −f (1) = − . (48.12)
k→0 4π~2
6) Is the relation
1
−Fl− (k)fl+ (k; r) + Fl+ (k)fl− (k; r)
φl = (48.13)
2ik
well defined for k ∈ C? Yes, it is, since for k ∈ C, φl , fl+ , fl− are still linearly
dependent, and we can still define φl , fl± in the same way, through
lim e±ikr fl± (k; r) = 1 , (48.14)
r→∞
but in this case, since k ∈ C, fl± (k; r) is not a pure phase at infinity, since |eikr | =
6 1.
F + (k)
7) If limr→0 lk2 is a constant for l even, to find the k → 0, r → ∞ behaviour
with k of φl (k; r) for even l we write
1
−Fl− (k)fl+ (k; r) + Fl+ (k)fl− (k; r) ,
φl (k; r) = (48.15)
2ik
and at r → ∞ fl± (k; r) ∼ e∓ikr , while for k → 0, Fl+ (k) ∼ k 2 × constant, and
Fl± (−k) = Fl∓ (k), so
Fl− (−k) Fl+ (k)
lim = lim = constant ≡ A , (48.16)
k→0 k2 k→0 k2
and for l even, Fl± (−k) = Fl (k), so finally then
eikr − e−ikr
φl (k; r) ∼ Ak = Ak sin kr. (48.17)
2i
49 Chapter 49
1) We use the Lippman-Schwinger equation for ψ, in the form of eq. 49.16 in the
text,
Z
2m ~0 0
fk (θ) = 2
d3~r0 e∓ik ·~r V (~r0 )ψ± (~k; ~r0 ) , (49.1)
4π~
to write an expression for σl . Indeed,
dσ
= |fk (θ)|2 , (49.2)
dΩ
P
and σtot = l≥0 σl , where
π
σl = (2l + 1)|e2iδl (k) − 1|2 = π(2l + 1)|al (k)|2
k2
Z 1 2
= π(2l + 1) d(cos θ)Pl (cos θ)fk (θ)
−1
Z 1 Z 2
2m ~0 0
= π(2l + 1) d(cos θ)Pl (cos θ) d3~r0 e∓ik ·~r V (~r0 )ψ± (~k; ~r0 ) .(49.3)
−1 4π~2
2) For the details of going from (49.19) to (49.20) in the text, from
2mik ∞ 0 02
Z
(1)
Rl (k; r) = jl (kr) − dr r jl (kr< )hl (kr >)V (r0 )Rl (k; r0 ) , (49.4)
~2 0
where r< , r> refers to the smaller and larger between r, r0 , we note that Rl (k 0 ; r)
(the left-hand side) is from the expansion of ψ in (2l + 1)Pl (cos θ) (see eq. (48.24)
in the text), and jl (kr) on the right-hand side is from the expansion in (2l +
~
1)Pl (cos θ) of eik·~r , and the second term on the right-hand side is from the expansion
ikr
of f (1) (~k, ~k 0 ) e r in l≥0 al (2l + 1)Pl (cos θ) (note that at r → ∞, thus for r0 < r,
P
(1) ikr (2) −ikr (1,2)
we have hl (kr) ∼ eikr i−l , hl (kr) ∼ e ikr il , and hl = jl ± inl ).
The integral on the right-hand side equals
Z r Z ∞
(1) (1)
dr0 r02 jl (kr0 )hl (kr)V (r0 )Rl (k; r0 ) + dr0 r02 jl (kr)hl (kr0 )V (r0 )Rl (k; r0 ) ,
0 r
(49.5)
and for r → ∞, the second integral can be neglected, finally arriving at
2m ∞ 0 02
Z
al = − 2 dr r jl (kr0 )V (r0 )Rl (k; r0 ). (49.6)
~ 0
q.e.d.
3) For V (r) = −λδ(r − a), we want to prove that Rl (k; r) is given by (49.31) and
the Jost functions by (49.32) in the text.
172
173 Chapter 49
From the Lippman-Schwinger equation for the radial Green’s functions in eq.
49.28,
Z
(0) (0)
g̃l (r; r0 ) = g̃l (r; r0 ) + dr00 r002 g̃l (r; r00 )V (r00 )g̃l (r00 ; r0 ) , (49.7)
we obtain
(0) (0)
g̃l (r; r0 ) = g̃l (r; r0 ) − λa2 g̃l (r; a)g̃l (a; r0 ) , (49.8)
so
(a;r 0 )
0 (0) (0) g̃l
g̃l (a; r ) = g̃l (a; r0 )−λa2 g̃l (a; a)g̃l (a; r0 ) ⇒ g̃l (a; r ) = 0
(0)
, (49.9)
1 + λa2 g̃l (a; a)
which can be put back in the Lippman-Schwinger equation to obtain (49.30) in the
text, namely
(0) (0)
(0) g̃l (r; a)g̃l (a; r0 )
g̃l (r; r0 ) = g̃l (r; r0 ) − λa2 (0)
. (49.10)
1 + λa2 g̃l (a; a)
Then, from (49.27) in the text,
Z
+(0)
Rl (k; r) = jl (kr) + dr0 r02 g̃l (E; r; r0 )V (r0 )Rl (k; r0 ) , (49.11)
where the left-hand side corresponds to g̃l and the first term on the right-hand side
(0)
corresponds to g̃l , then (49.30) in the text implies
(0)
g̃l (E; r; a)jl (kr)
Rl (k; r) = jl (kr) − λa2 +(0)
, (49.12)
1 + λa2 g̃l (E; a; a)
which is (49.31) in the text.
For the Jost solutions, φ∓
l (k; r) ∼ e
±ikr
at r → ∞, but
so we obtain
m2 k
X Z
(2l + 1)(Imal )Pl (1) = [(2π) 3
] dΩ|h~k|T̂ |~k 0 i|2 . (49.16)
2~2
l
5) For a potential with finite range and V = V0 > 0 for r < r0 and V = 0 for
r > r0 , with E > V0 , we write the solution
(
Cl jl (kr) + Dl nl (kr) , r > r0
REl (r) = , (49.20)
C̃l jl (k̃r) + D̃l nl (k̃r) , r < r0
2 2
k̃
where ~2m = E −V0 . Imposing regularity at r = 0, since jl (x) ∼ xl and yl (x) ∼ x−l ,
we must put to zero its coefficient, so D̃l = 0. Then at r = r0 we have the matching
conditions
j 0 (kr0 ) cos δl − n0l (kr0 ) sin δl j 0 (k̃r0 )
q0l = kr0 l = k̃r0 l . (49.21)
jl (kr0 ) cos δl − nl (kr0 ) sin δl jl (k̃r0 )
6) For the hard sphere, we have
jl (kr0 )
tan δl = , (49.22)
nl (kr0 )
so
π 2 4π 1
σl = (2l + 1) e2iδl (k) − 1 = 2 (2l + 1)
k 2 k 1 + tan21δl (k)
4π 1
= 2 (2l + 1) n 2 (kr ) . (49.23)
k 1 + 2l 0 jl (kr0 )
7) We consider the high energy limit, k → ∞, for σl of the hard sphere, and
obtain
jl (kr0 ) (l + 1)π
tan δl (k) = → cot kr0 − , (49.24)
nl (kr0 ) 2
so it oscillates! That means that, as a function of l, the contribution to
4π 1
σl = (2l + 1) (49.25)
k2 1 + tan21δl (k)
varies.
50 Chapter 50
i0l (k̃r0 )
q0l = k̃r0 , (50.6)
il (kr0 )
to be matched against the value at r > r0 .
2) We want to find the first correction to (50.30),
~2 κ2 ~2 1
Ibd.state = = , κ'− , (50.7)
2m 2ma2 a
between the binding energy of the bound state close to zero and the scattering
length. But
~2 κ2 ~2 ~2
1 2r0
κ= ⇒ Ibd.state = = 2 ' 1+ . (50.8)
r0 − a 2m 2ma2 1 − r0 2ma2 a
a
~2 k̃ 2 V0 3V0
=E− = , (50.9)
2m 2 2
175
176 Chapter 50
and kr0 1 means that the outside solution can again be rewritten as Be−kr , and
= C(r − a) (at low energy), so
~2 κ2 ~2
T = = . (50.10)
2m 2m(r0 − a)2
4) Extending from l = 0 to general l the analysis of the bound state for Sl (k),
δl (k), al (k), we have for a single bound state
k + iκ
Sl (k) ' − = e2iδl (k) , (50.11)
k − iκ
and
Sl (k) − 1 2 1
al (k) = '− = ⇒
2ik −κ − ik k cot δl (k) − ik
k cot δl (k) ' −κ. (50.12)
k + iκ ilπ F + (k)
Sl (k) = − e = e2iδl (k) = l− eilπ , (50.13)
k − iκ Fl (k)
and from the theorem stating that the zeroes of Fl+ (k) in the lower-half plane (Im
k < 0) are on iR (the imaginary axis) ↔ bound states, and are simple zeroes, so
Fl+ (k) = A(k + iκ) ⇒ Fl− (k) = Fl+ (−k) = −A(k − iκ). (50.14)
7) We want to prove
2 −4kn2
Nnl = , (50.15)
dFl+ (−iκ)
Fl− (−iκ) dκ
κ=κn
where, for physical bound states, the physical reduced radial function is
We have
1
−Fl− fl+ (k; r) + Fl+ (k)fl− (k; r) ,
φl = (50.17)
2ik
where fl± (k; r) ∼ e∓ikr for r → ∞, so
so for consistency we must have Fl+ (−iκn ) = 0 (so that there is no diverging eκn r
component). Then, for k 6= −iκn (κ 6= κn ),
d +
Fl− (iκ) = Fl+ (−iκ) = F (−iκ) , (50.21)
dκ l κ=κn
1) To write down the approximate form for Sl (k) with two resonances and 2 bound
states, we note that for bound states, near them we have
k + iκi ilπ
Sl (k) ' − e . (51.1)
k − iκi
On the other hand, for resonances of Sl (k), we saw that if kpole = k1 − ik2 , then
kzero = k1 + ik2 , which in turn means kzero = −k1 + ik2 , and kpole = −k1 − ik2 .
In general then, for the resonance,
k − k1 − ik2
Sl (k) ' . (51.2)
k − k1 + ik2
If the resonances and bound states are isolated, they don’t interfere, and then we
expect to have
(1) (1) (2) (2)
k + iκ1 k + iκ2 k − k1 − ik2 k − k1 − ik2
Sl (k) ' −eilπ , (51.3)
k − iκ1 k − iκ2 k − k (1) + ik (1) k − k (2) + ik (2)
1 2 1 2
(2) (1) (1) (2)
but, as we saw, the resonances come in pairs, so k2 = k2 , k1 = −k1 .
Another possibility is to have a sum of terms instead of the product, which will
also be an approximation (the correct result will be, of course, neither formula in
general)
2) If only σ1 is near the resonance, and in general
4π k22 4π (Γ/2)2
σl (k) = (2l + 1) 2 = (2l + 1) , (51.4)
k2 (k − k1 )2 + k2 k2 (E − E1 )2 + (Γ/2)2
then only for σ1 we have k ' k1 . Since, by assumption, near the resonance k2 k1 ,
and σl ≤ σ1 , then (if the other σl ’s are far from the resonance) σtot ' σ1 .
3) If Sl (k) has a single pole (”resonance”), but with k1 ∼ k2 , from (51.15) near
the pole
4π k22
σl (k) ' 2 (2l1 ) (51.5)
k (k − k1 )2 + k22
is still valid, and also
~2 2 ~2
Eres = (k1 − k22 ) , Γ/2 = 2k1 k2 . (51.6)
2m 2m
~2
Moreover, for k ' k1 , again E − E1 ' 2m 2k1 (k − k1 ), but in general
~2 2 ~2 2
Eres = (k1 − k22 ) = E1 − k . (51.7)
2m 2m 2
178
179 Chapter 51
~2 2
Eres = E1 − k 6= E1 . (51.11)
2m 2
4) If we have two resonances close to each other for σl , and at resonance E ' E1
(on the real line), the radial wave function is
eikr e−ikr
Rl (k; r) = C Sl (k) +
" r r #
(1) (2)
k − k1 − ik2 k − k1 − ik2 eikr e−ikr
'C (1) (2) r
+ , (51.12)
k − k1 + ik k − k2 + ik r
2 2
so
eikr e−ikr
χl (k; r) ' (1) (2)
+ (1) (2)
.
(k − k1 + ik2 )(k − k1 + ik2 ) (k − k1 − ik2 )(k − k1 − ik2 )
(51.13)
Consider k = k1 + δk, such that we have a wave packet
eiδk(r−v1 t)
Z
E1
ψ(r, t) = ei(k1 r− ~ t) dδk (1) (2)
(δk + ik2 )(δk + ik2 )
eiδk(r−v1 t)
Z
E1
−e−i(k1 r− ~ t) dδk (1) (2)
" (δk − ik2 )(δk − ik2 ) #
(1) (2)
E1 ek2 (r−v1 t) ek2 (r−v1 t)
= −2πiei(k1 r− ~ t) (2) (1)
+ (1) (2)
θ(v1 t − r)
i(k 2 − k 2 ) i(k2 − k 2 )
(1) (2)
" #
−i(k1 r−
E1
t) ek2 (r+v1 t)
ek2 (r+v1 t)
+2πie ~
(1) (2)
+ (2) (1)
θ(−v1 t − r).
i(k2 − k2 ) i(k2 − k2 )
(51.14)
5) In the case at the previous exercise, in the ”in” region, for t < −r/v1 , there
are 2 waves,
(1) (2)
k2 (r+v1 t)
E1
t) e − ek2 (r+v1 t)
ψ(r, t) = 2πie−i(k1 r− ~
(1) (2)
, (51.15)
i(k2 − k2 )
180 Chapter 51
in the intermediate region −r/v1 < t < r/v1 , again we have ψ = 0, and in the
”out” region, for t > r/v1 , again we have two waves,
(1) (2)
k2 (r−v1 t)
E1
t) e − ek 2 (r−v1 t)
ψ(r, t) = −2πiei(k1 r− ~
(2) (1)
. (51.16)
i(k2 − k1 )
6) For the Levinson theorem, nlb only means poles of Sl (k) on the imaginary line,
or also a bit off it? nlb is the number of energy levels = number of bound states =
number poles of Sl (k) on the imaginary line.
If the poles are not on the imaginary line, but a bit off it, k = iκ + δk,
eikr 1 e−ikr e−kr+iδkr 1 eκr−iδkr
ψ∼ − ∼ − . (51.17)
r Sl (k) r r Sl (iκ + δk) r
If the bound state is a pole at iκ + δk, then
e−κr+iδkr
ψ∼ , (51.18)
r
so we have an extra phase, otherwise it seems OK, which would seem to indicate
the pole could be slightly off as well. However, in practice, then REl (r) would not
be real anymore, so the factor sin(kr − lπ/2 + δl ) will get changed. So, in fact, we
cannot have the pole a bit off.
7) If Sl (k) has a single resonance very close to R for all l, by considering l ∈ C,
S(l; E) has a pole = zero of B(l; E) = Fl− (k), so the Regge trajectory collapses to
a point on the real line (all l’s, which should be = αi (E), now are at a single point).
52 Chapter 52
1) For the potential V (x) = V0 /(x2 +a2 ), scattering with an energy 0 < E < V0 /a2 ,
the turning points are
r
V0
x1,2 = ± − a2 . (52.1)
E
Considering a wave coming in from the left, for x < x1 we have the transition
rule at x1
√ " Z s #
1 x 0
2 V0 π
i1/4 sin ~ dx 2m E − 2 −
h x + a2 4
2m E − x2V+a
0
2
x1
" s #
1 x 0
Z
1 V0
→h i1/4 exp − ~ dx 2m −E . (52.2)
V0 x1 x2 + a2
2m x2 +a2 − E
Then, at x2 , we have
" s #
Z x2
1 1 0 V0
i1/4 A exp + ~ dx 2m −E
h x + a2
2
2m x2V+a
0
2 − E x
√ " Z s #
1 x 0
A 2 V0 π
→h i1/4 sin ~ dx 2m E − 2 + , (52.3)
x + a2 4
2m E − x2V+a0
2
x2
where
" s #
Z x2
1 0 V0
A = exp − dx 2m −E . (52.4)
~ x1 x2 + a2
2) For the 3d problem with V (r) = V0 /(r2 + a2 ) and 0 < V0 < V0 /a2 , the
transition point is
r
V0
r0 = − a2 , (52.5)
E
and the wave function of angular momentum l is
√
kr
ψ = Ylm (θ, φ) W , (52.6)
r
√
where krW = χ(r), and in the WKB approximation, for r < r0 ,
Z x0
1
WWKB (x ≡ ln kr) = p exp − dx0 κ(x0 ) , (52.7)
κ(x) x
181
182 Chapter 52
where
2 2mr V0
k (x) = 2 E− − (l + 1/2)2 , k 2 (x) = −κ2 (x) , (52.8)
~ r2 + a2
and for r > r0 ,
Z x
2 π
WWKB (x = − ln kr) = p sin dx0 k(x0 ) + . (52.9)
κ(x) x0 4
and
4π 2
X
σl = (2l + 1) sin δ l , σ tot = σl . (52.11)
k2
l
If r0 a (so if V0 /E a2 ), we have
"r # "r #
Z ∞ Z ∞ ∞
0 A 0 A 2 /k 2 A2 1 A2
dr k 2 − 02 − k = k dr 1− 02
−1 ' 2 0 =− ,
r0 r r0 r 2k r r0 2k 2 r0
(52.12)
where
2mV0
A≡ − (l + 1/2)2 . (52.13)
~2
4) For V (r) = V0 /(r2 + a2 ), in the eikonal approximation,
Z +∞
mV0 +∞
Z
m p dz
∆(b) = − 2 dzV ( b2 + z 2 ) = − 2
2~ k −∞ 2~ k −∞ z + a2 + b2
2
mV0 π
=− √ . (52.14)
2~2 k a2 + b2
If V = 0 for r ≥ r0 , for b < r0 we have
Z +√r02 −b2
m V0
∆(b) = − 2 dz 2
2~ k −√r02 −b2 z + a2 + b2
p
mV0 1 r2 − b2
=− 2 √ arctan √ 0 . (52.15)
~ k a2 + b2 a2 + b2
If r0 is large, then in the eikonal approximation,
Z r0 h i
f (+) (θ) = ik db bJ0 (kbθ) e2i∆(b) − 1 , (52.16)
0
√
r02 −b2 π
and, for r0 a, b, we obtain arctan √
a2 +b2
' 2, so
mV0 π
∆(b) ' − √ , (52.17)
2~2 k a2 + b2
183 Chapter 52
so
Z r0 →∞
(+) πiV0
f (θ) = ik db bJ0 (kbθ) exp − √ −1
0Z ~2 k a2 + b2
∞ 2
i πimV0 θ/~
= 2 dz zJ0 (z) exp − √ −1 . (52.18)
kθ 0 a2 k 2 θ2 + z 2
−µr
5) For the Yukawa potential V (r) = V0 er , in the eikonal approximation, we
can use the formula for a finite range r0 ∼ 1/µ for the cross section,
Z r0
σtot = 8π db b sin2 ∆(b) , (52.19)
0
where
√
mV0 +∞ e−µ b +z
Z 2 2
∆(b) = − 2 dz √
2~ k −∞ b2 √
+ z2
mV0 +1/µ e−µ b +z
Z 2 2
mV0 2
'− 2 dz √ ∼− 2 , (52.20)
2~ k −1/µ 2
b +z 2 2~ k µb
so that
2 Z ~2 k
Z 1/µ
1 mV0
2 8π mV0 mV0 1
σtot ∼ 8π db b sin = 2 dx x sin2 , (52.21)
0 µb ~2 k µ ~2 k 0 x
where the integral is almost a number (if ~2 k/mV0 → ∞).
6) For a Coulomb potential V (r) = A/r, but with V = 0 for r ≥ r0 , in the
eikonal approximation
Z r0
eik
σtot ' 8π db b sin2 ∆(b) , (52.22)
0
and
√ √
Z + r02 −b2 + r02 −b2
mA dz mA 1 z
∆(b) = − 2 √ √ =− 2 arcsinh √
2~ k − r02 −b2
2
b +z 2 2~ k b b − r02 −b2
r
mA r02
=− arcsinh −1, (52.23)
~2 kb b2
and for r0 /b 1, the arcsinh becomes ' ln 2rb0 , so that
Z r0 Z r0
mA 2r0
eik
σtot = db b sin2 ∆(b) ' db b sin2 ln , (52.24)
0 0 ~2 kb b
and this is also divergent as r0 → ∞, as is the classical Rutherford formula.
7) We want to check if there are any resonance for Coulomb scattering, and the
analytical properties of Sl (k). We have
Sl (k) − 1
al (k) =
X 2i
f (θ) = (2l + 1)al (k)Pl (cos θ)
l
184 Chapter 52
1) Mapping the black disk eikonal into the general δl (k), via δl (k) ↔ ∆(b)|b=l/k ,
we remember that
Then we have
kr0
inel π X π 2 2
σtot (k) = (2l + 1) = kr 0 (kr0 + 2) = πr0 1 + . (53.3)
k2 k2 kr0
l=0
2) We have
2
dσ ikr0 k 2 r02 2
= |f (θ, r0 )|2 = − J1 (−qr0 ) = J (−qr0 ). (53.4)
dΩ q q2 1
Ze20 4π
Z
1 3 ∗ −i~
q ·~
r1
= −δ 2,1 + d ~
r ψ
1 n=2,l=0 1 (~
r )e ψ (~
r
n=1,l=0 1 )
q2 Z
2 Z ∞ Z π −iqr1 cos θ
e 4π e ∗
= 0 r12 dr1 2π sin θdθ R2,0 (r)R1,0 (r)
q Z0 0 4π
2 ∞
e 4π ∗
= 02 r1 dr1 sin(qr1 )R2,0 (r1 )R1,0 (r1 ). (53.8)
q 0
~
X eik0 r
~ +
= eik·~r un0 (~r)gB (R) fnn0 (~k 0 , ~k)un (~r)gB (R)
~ (53.18)
,
n
r
1) Yes, the wave function (or something like it, the quantum field, in the relativistic
case) satisfies the Klein-Gordon equation. The (relativistic) Dirac equation is
∂
i~ ψ = i~cγ 0 γ i ∂i ψ + imc2 γ 0 ψ. (54.1)
∂t
Squaring it, we get the KG equation:
~ 2 ψ − (mc2 )2 ψ = 0.
[~2 c2 ∂µ2 − (mc2 )2 ]ψ = −~2 ∂t2 ψ + ~2 c2 ∇ (54.2)
But then: taking the square root to go back, we get ±mc2 : the rest energy con-
sidered with arbitrary sign (but negative energy states are ”occupied”, giving the
”Dirac sea”, according to Dirac’s interpretation).
2) The Dirac equation in 1+1 dimensions is formally the same in terms of the
arbitrary gamma matrices,
~ ∂
γ0 ψ + ~γ i ∂i ψ + mcψ = 0 , (54.3)
c ∂t
or
(γ µ ∂µ + m)ψ = 0 (54.4)
for ~ = c = 1, where
{γ µ , γ ν } = 2η µν . (54.5)
{σ i , σ j } = 2δ ij , (54.6)
σ 1 = γ 1 , σ 2 = iγ 0 ⇒ {γ µ , γ ν } = 2η µν . (54.7)
3) We have
γ5 = iγ 0 γ 1 γ 2 γ 3 ⇒
γ52 = −γ 0 γ 1 γ 2 γ 3 γ 0 γ 1 γ 2 γ 3 = +γ 0 γ 1 γ 2 γ 0 γ 1 γ 2
= γ 0 γ 1 γ 0 γ 1 = −(γ 0 )2 = +1
{γ5 , γ3 } = i(γ 0 γ 1 γ 2 (γ 3 )2 + γ 3 γ 0 γ 1 γ 2 γ 3 )
= i(γ 0 γ 1 γ 2 − γ 0 γ 1 γ 2 ) = 0 , (54.8)
188
189 Chapter 54
where we have used the definition of the Clifford algebra to anticommute matrices
and square them, and similarly
{γ5 , γ2 } = {γ5 , γ1 } = {γ5 , γ0 } = 0 , (54.9)
so that finally we have the relations missing to form the 5-dimensional Clifford
algebra,
[{γ5 , γM } = 2η5M , M = (5, µ) , {γµ , γν } = 2ηµν ] ⇒ {γM , γN } = 2ηM N . (54.10)
4) To couple the KG equation to electromagnetism, we start with the Dirac
equation coupled to electromagnetism (the relativistic form),
~
γµ ∂µ − qAµ + mc ψ = 0 , (54.11)
i
and since the square of the Dirac equation gives the KG equation, we square the
Dirac equation (multiplied by γ 0 ) coupled to electromagnetism, to give
0 µ ~ 0 0 ν ~ 0
γ γ ∂µ − qAµ + mcγ γ γ ∂ν − qAν + mcγ ψ = 0 ⇒
" i i #
2
~ ~
∂µ ∂ µ + q 2 Aµ Aµ + (mc)2 − q 2∂ µ Aµ ψ = 0. (54.12)
i i
5) In the shell model for the nucleus, with the nucleon potential being Vn (r) =
−µr
−V0 e r , the relativistic corrections to the Hamiltonian for the nucleon are found
just like the corrections to the Hamiltonian for the Hydrogenoid atom, namely the
relativistic corrections to the Schrödinger equation are
p2 )2
(~ 1 ~ ~ 1 dVn ~2
Hn0 = − 3 2
+ L·S + ∆Vn , (54.13)
8m c 2m r dr 8m2 c2
−µr
applied for Vn = −V0 e r , so that
e−µr µ2
1 dVn V0 −µr 1 −µr 3
= 2e µ+ , ∆Vn = −V0 ∆ = −V0 e 4πδ (r) + .
r dr r r r r
(54.14)
6) We check the missing steps in the relativistic corrections to the Hydrogenoid
atom. We have
p2 )2
(~ 1
hH1 inljmj = h− p2 )2 inljmj ,
inljmj = − 3 2 h(~ (54.15)
8m3 c2 8m c
and
2 2
2 ~ ∂ 2 ∂ l(l + 1)
p~ = + − ,
i ∂r2 r ∂r r2
4 4 4
4m c (αZ) 1 3
p2 )2 inljmj =
h(~ 3
− , (54.16)
n l + 1/2 4n
and
1 Z2
h i nl = , (54.17)
r2 n3 a20 (l + 1/2)
190 Chapter 54
(Zα)4 mc2
+1
= 3 , (54.25)
4n (l + 1)(l + 1/2) −(1 + 1/l)
which matches what we wanted to prove, at large l (+1/(−1 − 1/l) vs ±1).
7) For the relativistic corrections to the Hydrogenoid atom in magnetic field, we
want the corrections to
(~ ~ 2
p − q A) ~q
− ~ ,
~σ · B (54.26)
2m 2m
and in a constant magnetic field, A ~=B ~ × ~r/2, so the interaction Hamiltonian has
the usual linear term in B
q ~ ~ q ~ ~ q ~ ~ ~ ,
Hint = − B·L− B·S =− B · (L + 2S) (54.27)
2m 2m 2m
so the calculation follows the previous one (in the text), with the interaction given
by the Landé g-factor,
qB
∆E = −g mj . (54.28)
2m
191 Chapter 54
~ of the corrections
Then the first relativistic correction comes from coupling to A
in the text, which gives an extra contribution only in H1 , where we now have
* +
~ 2 ]2
p − q A)
[(~
− , (54.29)
8m3 c2
~= ~ r
B×~
out of which, from A 2 , and after the rearrangement, we get
~ · L)
(q B ~ 2 q 2 B 2 ~2 m2l
− 3 2
→− , (54.30)
8m c 8m3 c2
so the contribution to the energy is quadratic in B.
55 Chapter 55
5) For the case at the previous exercise, we write T (1) in the occupation number
basis as (N = 5 for 5 bosons)
5 r
X
0 n0 !...n3 ! 0
T(1) = hi|T̂(1) |i i Cn0 ...ni −1...ni0 +1...n3 (t) , (55.8)
0
5!
i,i =1
192
193 Chapter 55
where hφEi |T̂i |φEi0 i ∝ δEi Ei0 is diagonal, and leaves the Slater determinant invariant.
56 Chapter 56
is also diagonalized, so
eαb |βi = eαβ |βi. (56.4)
2) The Hamiltonian
is interpreted physically as follows: the first term is the kinetic energy Ĥ0 , which
has no interactions. Then b†i bi+1 means hopping from site i + 1 to site i, and b†j bj−1
hopping from j − 1 to j, so the interaction is between 2 hoppings.
3) For a system with 7 one-particle states and 3 bosons in it, the Fock space,
where we commuted the operator ai over (1 − αa† ), used a|αi = α|αi and α2 = 0
and then recomposed the state.
Finally then, we have
P X X
e( i ai )α |β1 ...βn i = (1 − α ai )|β1 ...βn i = (1 − α βi )|β1 ...βn i
P i i
= e( i βi )α
|β1 ...βn i. (56.10)
5) For the details of the fermionic Schrödinger equation on the occupation num-
ber states, we note that (56.14) in the text is the same for fermions, except that
|ni − 2i ≡ |0i, so no last term. But also (56.18) in the text follows, since there are
an even number of b’s and b† ’s. So (56.19) in the text is the same, hence (56.20) in
the text gives
X † 1 X † †
Ĥ = ai hi|T̂ |i0 iai + ai aj hij|V̂(2) |i0 j 0 iai0 aj 0 . (56.11)
0
2 0 0
i,i iji j
6) For the potential V (~ri − ~rj ) = V0 δ(~ri − ~rj ), the field operator is
Z Z Z
1
ψ̂σ† (~r)ψ̂σ† 0 (~r0 )V̂ (~r, ~r0 )ψ̂σ0 (~r0 )ψ̂σ (~r)
X X
†
Ĥ = d ~r3
ψ̂σ (~r)T̂ (~r)ψ̂σ (~r) + d ~r d3~r0
3
σ
2 0
Z Z σ,σ
V0
ψ̂σ (~r)ψ̂σ† 0 (~r)ψ̂σ0 (~r)ψ̂σ (~r)
X X
3 † 3 †
= d ~r ψ̂σ (~r)T̂ (~r)ψ̂σ (~r) + d ~r
σ
2 0
σ,σ
X † V0 † †
= ti ci ci + ai aj aj ai . (56.12)
i
2
and the Feynman diagram is for an incoming fermion with momentum ~k 0 and σ 0
emitting a photon with momentum ~q that is absorbed by another fermion with
momentum ~k and σ, so the final state has fermions with momentum ~k 0 − ~q and σ 0
and one with momentum ~k + ~q and σ.
57 Chapter 57
|δψi = δb†i1 b†i2 ...b†iN |0i + b†i1 δb†i2 ...b†iN |0i + ...b†i1 ...b†iN −1 δb†iN |0i , (57.3)
and we have
δb†i = −i Kji b†j .
X
(57.4)
j
If δb†i contains other b†j (so that Kji 6= 0 for i 6= j), then hδψ|ψi = 0, so we obtain
hδψ|Ĥ|ψi = 0 , (57.5)
but
hδψ|Ĥ(2) |ψi = h0|...bi ...bj ...|Ĥ(2) |...b†i0 ...b†j 0 ...|0i = hij|Ĥ(2) |i0 j 0 i. (57.6)
That means that terms with one index on the left equal to one index on the right
must vanish, and the equality to zero will give the Hartree-Fock equation. We have
the relevant terms in Ĥ(2)
1X 1X
hij|V̂(2) |ij 0 ib†i b†j bi bj 0 + hij|V̂(2) |i0 iib†i b†j bi0 bi
2 0 2 0
ijj iji
1X 0 † † 1X
+ hij|V̂(2) |jj ibi bj bj bj 0 + hij|V̂(2) |i0 jib†i b†j bi0 bj
2 0 2 0
ijj
X iji
= hik|V̂(2) |i ki + hik|V̂(2) |ki i b†i b†k bk bi0
0 0
i,k,i0
196
197 Chapter 57
!
b†i b†k bk hik|V̂(2) |i0 ki b†k bk hik|V̂(2) |ki0 i
X X X
= + bi0
i,i0 k k
" #
b†i
X X X
0 0
= Nk hik|V̂(2) |i ki + Nk hik|V̂(2) |ki i bi0
i,i0 k k
b†i hi|V̂H−F |i0 ibi0 .
X
≡ (57.7)
i,i0
Then the terms with one only one b†i changed must vanish, giving the equation
where
X e20
ρ(~r) = |φk (~r)|2 , V(2) (~r, ~r0 ) = , (57.11)
|~r − ~r0 |
k,occ.
and for He in the ground state there are 2 electrons in the 100 state, so
X Z e20 |φ100 (r0 )|2
Z
VH,He ground state = d3~r0 |φ100 (|~r0 |)|2 0
= 2e 2
0 d3~r0
|~r − ~r | |~r − ~r0 |
k=1,2
4r 0
e− a0
3 Z 1 Z ∞
4e20 2 02 0
= d cos θ r dr q
r a0 0 2 0
−1 0 1 + rr − 2 rr cos θ
3 Z ∞
r0 r0
2 2 0 0 − a0
4r 0
= 4e0 r dr e 1+ − 1− , (57.12)
a0 0 r r
where we have used that for the Helium atom (Hydrogenoid with Z = 2) we have
3/2 R +1 dx
φ100 (r) = a20 2e−2r/a0 √14π and did the integral over cos θ, using −1 √a−bx =
2
√ √
b( a + b − a − b).
At large r (r a0 ), we can take r > r0 in the integral (since the integral is cut
0
off by e−4r /a0 ), and obtain
3 Z ∞
4e20 2 4r 0 e2
VH ' r02 dr0 e− a0 = 0 , (57.13)
r a0 0 r
198 Chapter 57
which is the Coulomb potential of a single positive charge, like for the He nucleus
screened completely by the other electron. At intermediate r the result can be
written in terms of incomplete gamma functions, or otherwise be left as the integral
above.
5) For Lithium, Z = 3 and the electrons are in the configuration 1s2 , 2s1 . Then
then 2-particle exchange term in the Hartree-Fock approximation is
(2)
X φ∗ (~r0 )φk (~r) e2 1
VH−F (~r, ~r0 ) = e20 k
0
= 0 [2R10 (r0 )R10 (r) + R20 (r0 )R20 (r)]
|~r − ~r | 4π |~r − ~r0 |
k,occ.
3
e20 3r0
1 3 − a3 (r+r 0 ) 1 − 2a3 (r+r0 ) 3r
= 8e 0 + e 0 2 − 2 − ,
4π |~r − ~r0 | a0 8 a0 a0
(57.14)
where we have used
3/2 3/2
Z − Zr Z Zr Zr
R10 (r) = 2e a0
, R20 (r) = 2− e− 2a0 . (57.15)
a0 2a0 a0
6) The Hartree-Fock equation for the 2-point Green’s function is
(2) (2) (2)
G(2) ' G0 + G0 · VH−F · G(2) , (57.16)
so explicitly, in the (~r, σ) representation, is
(0)
GσσZ0 (~r, t; ~r0Z, t0 ) ' ZGσσ0 (~r, t; ~r0 , t0 )+
(0) (2)
X
d3~r00 dt00 d3~r000 Gσσ00 (~r, t; ~r00 , t00 )VH−F (~r00 , ~r000 )Gσ00 σ0 (~r000 , t00 ; ~r0 , t0 ).
σ 00
(57.17)
7) The Dyson, or version of the Bethe-Salpeter equation, is
(2) (2) (4) (2) (2)
G(4) = G0,1 · G0,2 + KB−S · G0,1 G0,2 · G(4) , (57.18)
and its explicit form in the the (~r, σ) representation, is
G(4) r1Z, t1 , ~r2 , Zt2 ; ~r3 , t3Z, ~r4 , t4Z) = GZ(0)
σ1 σ2 ,σ3 σ4 (~ r1 ,Zt1 ; r~3 , t3Z)G(0)
σ1 σ3 (~ σ2 σZ
4
(~r2 , t2 ; ~r4 , t4 )
X
+ d3~r10 d3~r20 dt01 dt02 d3~r100 d3~r200 dt001 dt002 ×
σ10 ,σ20 ,σ100 ,σ200
(4)
×KB−S σ1 σ2 ,σ0 σ0 (~r1 , t1 , ~r2 , t2 ; ~r10 , t01 ; ~r20 , t02 )Gσ10 σ100 (~r10 , t01 ; ~r100 , t001 )×
1 2
×Gσ20 σ”2 (~r20 , t02 ; ~r200 , t002 )Gσ100 σ200 ,σ3 σ4 (~r100 , t001 , ~r200 , t002 ; ~r3 , t3 , ~r4 , t4 ). (57.19)
58 Chapter 58
1) For an example of the fact that the wave function changes for a non-Abelian
Berry phase, yet it is consistent, consider the following: If the states are fully de-
generate, there is no observable way to discern between them, so changing from one
(ab)
to the other by Peiγ is OK.
2) For the steps leading to (58.12), starting with the Hamiltonian
2
X p~i − q A(~~ ri ) X X
H= + V(~ri − ~rj ) + qA0 (~ri ) , (58.1)
i
2m i<j i
~ i,
we note that, since p~i = ~i ∇
U −1 p~i − q A(~ ~ ri ) + U −1 ~ ∇
~ ri ) U = p~i − q A(~ ~ iU
i
~ ri ) − ~θ
X
= p~i − q A(~ ~ i α(~ri − ~rj ) ,
∇ (58.3)
π i<j
and in 2 spatial dimensions, since arg(~x) = arctan y/x, we have ((x1 , x2 ) ≡ (x, y))
xj
∂i arg(~x) = −ij ⇒
|~x|2
(~ri − ~rj )b ẑ × (~ri − ~rj )
∇ai α(~ri − ~rj ) = azb = , (58.4)
|~ri − ~rj | |~ri − ~rj |2
where ẑ is the unit vector perpendicular to the 2 dimensional plane.
Then we finally have
X ẑ × (~ri − ~rj )
U −1 p~i − q A(~ ~ ri ) − ~θ
~ ri ) U = p~i − q A(~ . (58.5)
π i<j |~ri − ~rj |2
~ so what we obtain is
But, in general, we have the matrix relation ~j = σ · E,
(considering also what happens when we exchange the 1 and 2 indices)
k ij
σ ij = , (58.8)
2π
in other words, the integer quantum Hall effect (IQHE).
4) The Chern-Simons action is, in the Lorentz invariant form,
Z
k
SCS = d3 x ijk Ai ∂j Ak . (58.9)
B=∂M 4π
But by the Stokes theorem for general form integration (noting that the integrand
is a 3-form), this is equal to
Z
k
SCS = d4 x µνρσ (∂µ Aν )(∂ρ Aσ )
M Z 4π
k
= d4 xµνρσ Fµν Fρσ , (58.10)
16π M
where we have used that µνρσ ∂µ ∂ρ Aσ = 0, when we acted with the extra (anti-
symmetric) derivative on the CS integrand.
But note that the action is gauge invariant, since it is written only in terms of the
field strength Fµν , and it is topological, since there was no metric needed in order
√
to write it (not even the usual invariant integration measure d4 x −g), we used
R
the Levi-Civitta symbol µνρσ instead. That means that the integral is invariant
under small perturbations of the metric, so it is proportional to an integer, i.e., it
is topological.
5) We didn’t consider the equation of motion for Aµ when considering the CS
action or anyons because Aµ is an external source, so appears only linearly in the
full action. Then, when integrating out aµ , we obtain the CS action in (58.43) in
the text,
Z
0 1 1
Seff = d3 xijk Ai ∂j Ak , (58.11)
r 4π
which is quadratic, but contains no propagating degrees of freedom. In fact, it is
a response action, meaning that it defines the response of the system with respect
to the external source. That means that we don’t consider its equation of motion,
since Aµ lacks the dynamical kinetic term.
6) For FQHE we considered Fµν ∝ δ 2 (~r) unphysical because of flux conservation,
but then fµν ∝ δ 2 (~r) is OK, since the statistical gauge field is only defined in 2+1
dimensions. There is no physical 4th direction for it, like there is for Fµν .
7) The Moore-Read wave function is
Y
ψMoore−Read ∝ (zi − zj )m , (58.12)
i<j