(Michele Cini) Topics and Methods in Condensed Mat
(Michele Cini) Topics and Methods in Condensed Mat
(Michele Cini) Topics and Methods in Condensed Mat
Michele Cini
123
57/3180/YL
543210
Preface
This book is structured as a course for students with a good knowledge of basic Quantum Mechanics who want to specialize in Condensed Matter Theory.
Indeed it takes the reader to advanced levels in several topics, but there is an
obvious trick: no teacher can hope to cover all this material in one semester
without facing a serious and rightful student revolt. I am oering alternative
advanced topics that I have been developing in dierent years of teaching and
updated to the present time. Yet, I had to remove much important material
which is a condicio sine qua non for condensed matter theorists; the most serious sacrice (wisely urged by the Editor) was the removal of all the chapters
in Relativistic Quantum Mechanics and QED that I have been teaching for
many years in Atomic Physics courses but would have made the size of this
work unacceptable as one book. Such topics are covered in excellent Springer
books, like the series by Greiner. Even so, there is no encyclopedic attitude,
or attempt to cover all the hottest advanced topics, but I privileged those
arguments where I did some work and the methods that I used most often
or where I gave some contribution in my long research activity. The central
part of the book is devoted to the theory and applications of symmetry and
Greens functions. In the interest of the class, I present them is such a way
that one can easily separate an introductory part, that might be of interest
to a broader audience including experimentalists, and a more advanced and
demanding part.
Nobody really knows how knowledge grows, particularly in theoretical
physics (a rather special environment to study) but my impression is that
VIII
Preface
it grows by a vivid (if often not very precise initially) understanding of particular problems, that are later combined in more powerful units and made
more exact in the process. More complex phenomena require more mathematical ingenuity, new concepts arise, deep unexpected similarities between
diverse problems are discovered. By such processes in which mathematics and
physics merge into the reasoning process and are a source of inspiration for
each other, one can sometimes predict new facts that are later veried experimentally. In condensed matter physics, predicting new experimental features
is not reserved to geniuses, and one can really hope to achieve such results,
and although the success is not uncommon, it is certainly rewarding. I believe
that there is still lot of space for pioneers and while computing is important
in a quantitative science like physics most real discoveries will continue to
come from intuition and original thought.
Most times it is experiment that brings some unexpected and surprising
result. Superconductivity is a typical example: common sense would have
predicted that the resistance will unavoidably be present as a part of the imperfection of reality. This is the magic of Quantum Mechanics, that continues
to make stunning yet tangible reality by the eect of phases. The role of Berry
phases, ux quantization an the like is discussed mainly in Chapters 16 and
17. There are many phenomena but if we wish to go from phenomenology to
principles progress is conditioned by methods. A large variety of methods have
been generated by many theoreticians. Some are time honored, but enlighten
new problems in new ways: to make practical predictions in the presence of
symmetry we take advantage of abstract Group theory. However, important
new methods continue to be invented, and there is no sign that the gold mine
of ingenuity is exhausted. On the other hand, while an increasing number
of problems are successfully dealt with by the existing codes and computational methods, the ingenuity continues to be required, since new interesting
materials and processes are discovered which require fresh modeling.
Some of the subject matter is included because in my carrier I happened
to work in the subject or to make extensive use of the methods; in this way
I suppose I can provide in some measure a rst-hand presentation of some
topics. In particular, Chapters 6, 15, 13, 14 and 17 are the most original in the
sense of presenting results from my own research; elsewhere I tried to present
mature results in a fresh and attractive way. At least, I hope I will transmit
some of the fun I have to work using such results. The discussion of general
topics like Group theory or Feynman diagrams, the Keldysh theory, which
are discussed in many textbooks, but are traditionally considered rather hard
aims to help the reader by many examples and by direct procedures and
intuitive arguments. I tend to prove everything and therefore the reader will
nd sentences like it can be shown that ... only seldom and for relatively
unimportant side matters. Indeed, in most cases when something is really
understood, there is little diculty in nding a proof; on the other hand,
the converse is also true: having no clear justication often means having no
Preface
IX
real grasp or lacking mastery of the use of the results. However in several
cases, lengthy formal proofs are readily found in the literature and there is
no strong reason to reproduce them here: for instance, Wicks theorem and
the linked cluster theorem fall in this category. Then I prefer to give my own
intuitive arguments, that reect the way I visualize the result for myself, and
reference the literature where the canonical proofs are published.
Finally, I hope the readers will understand that this work costed me a
considerable eort and at some point I had to force the writing process to
converge. It is always possible to improve the book in many ways, but alas at
some point the writing must stop. It was this appeal to the readers understanding, besides the common interest in the Nature of Things, that prompted
to me the above quotation from Titus Lucretius Carus.
It is a pleasure to thank Professor Giancarlo Rossi, Doctor Gianluca Stefanucci and Doctor Yassen Stanev, all at the Physics Department of Rome
Tor Vergata University, and Doctor Claudio Verdozzi, currently at Lund, for
reading and discussing part of the manuscript and giving useful advice and
encouragement.
Michele Cini
Contents
11
13
15
16
18
18
19
19
21
21
23
25
26
29
29
29
1
1
2
4
5
5
7
9
35
37
38
38
40
XII
Contents
42
43
46
46
46
48
50
55
55
55
55
57
58
59
61
66
70
70
71
73
75
76
76
78
79
81
81
81
82
82
83
86
89
90
95
95
98
100
101
105
106
107
Contents
XIII
109
109
111
111
113
115
118
122
123
127
129
130
130
133
133
135
135
140
143
144
145
149
152
154
155
156
157
157
157
158
160
160
163
169
173
174
177
179
XIV
Contents
181
181
183
184
185
186
188
188
189
193
194
196
197
197
199
201
204
207
207
208
208
217
217
219
221
224
225
227
229
230
231
232
235
238
210
212
213
216
Contents
XV
241
242
245
245
247
248
249
251
253
254
257
257
259
263
264
264
267
270
272
272
274
275
276
277
278
279
279
282
283
13 Non-Equilibrium Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
13.1 Time-Dependent Probes and Nonlinear Response . . . . . . . . . . .
13.2 Kadano-Baym and Keldysh Methods . . . . . . . . . . . . . . . . . . . .
13.3 Complex-Time Integrals by Langreths Technique . . . . . . . . . . .
13.3.1 Finite temperatures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
13.4 Keldysh-Dyson Equation on the Contour . . . . . . . . . . . . . . . . . .
13.5 Evolution on Keldysh Contour . . . . . . . . . . . . . . . . . . . . . . . . . . .
13.5.1 Contour Evolution of Bosons . . . . . . . . . . . . . . . . . . . . . .
13.6 Selected Applications of the Keldysh Formalism . . . . . . . . . . . .
13.6.1 Atom-Surface Scattering . . . . . . . . . . . . . . . . . . . . . . . . . .
13.6.2 Quantum Transport . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
285
285
286
288
290
291
294
298
298
299
304
309
11.5
11.6
11.7
11.8
11.9
12
XVI
Contents
313
313
313
315
316
322
323
326
329
332
333
333
335
335
336
342
343
343
346
351
355
355
356
356
357
358
362
367
369
373
373
375
378
381
Contents
XVII
18 Algebraic Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
18.1 Lieb Theorems on the Half Filled Hubbard Model . . . . . . . . . .
18.2 Bethe Ansatz for the Heisenberg Chain . . . . . . . . . . . . . . . . . . .
18.3 Bethe Ansatz for Interacting Fermions 1 Dimension . . . . . . . . .
18.3.1 -Function Interaction . . . . . . . . . . . . . . . . . . . . . . . . . . . .
18.3.2 The Hubbard Model in 1d . . . . . . . . . . . . . . . . . . . . . . . . .
18.3.3 The Periodic Boundary Conditions . . . . . . . . . . . . . . . . .
18.3.4 Spin Chain Analogy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
383
383
387
392
392
396
401
403
406
Part VI Appendices
19 Appendix 1: Zero-point Energy in a Pillbox . . . . . . . . . . . . . . 409
20 Appendix II-Character Tables . . . . . . . . . . . . . . . . . . . . . . . . . . . . 411
21 Proof of the Wigner-Eckart Theorem . . . . . . . . . . . . . . . . . . . . . 413
Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 415
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 431
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 439
Part I
(1.1)
The electrons are identical, but this does not prevent us from labeling them;
rather it imposes that the wave function changes sign for each exchange of labels.
2
in words,P sends 1 2 and then Q sends 2 2, and in the same way 2
1 3 and 3 3 1.
1
()P P
N!
(1.2)
N
k
1
vk =
()P vQ1 (1)vQ2 (2) . . . vQN (N ),
N! Q
(1.3)
or, equivalently,
v1 (1) v2 (1)
1
v1 (2) v2 (2)
(1, 2, . . . N ) =
...
...
N!
v1 (N ) v2 (N )
. . . vN (1)
. . . vN (2)
.
... ...
. . . vN (N )
(1.4)
Note that the transposed matrix is equivalent since the determinant is the
same. Exchanging two rows, that is, two electrons, one gets a - sign. A permutation of the electrons is equivalent to the inverse permutation of the
spinorbitals, and P (1, 2, . . . N ) = (P1 , P2 , . . . PN ) = ()P (1, 2, . . . N ).
The set of all determinants is complete provided an arbitrary order is
xed for the one-electron states (otherwise the set is overcomplete).
one electron and then consider the same
Suppose we solve hwi = i wi for
N
problem with N electrons and H = i h(i). This problem is no harder than
for a single electron, and the N body Schr
odinger equation is solved by (1.4)
with energy eigenvalue 1 + 2 + . . . + N .
1.1.1 Many-electron Matrix Elements
The matrix element |F | of operators F (1, 2, . . . N ) between determinantal states, when we expand the determinants, means a sum of (N !)2 terms.
This grows disastrously with N ; however there are simple rules to calculate
such matrix elements. Let (1, . . . N ) be a determinant made by N spinorbitals u1 , u2 . . . uN ; taken out of the orthogonal set {wi } (they may be same
as in , in which case we are dealing with expectation values). One can
readily observe these rules by working out a 2 2 example, while the proof
requires a trick which is explained in Sect. 1.1.2. The simplest case is f = 1,
and the rule is: the overlap between determinants is the determinant of the
matrix with elements the one-electron overlaps ui |vj :
| = Det [{ui |vj }] .
(1.5)
This useful result holds even if u and v spinorbitals are taken from dierent
sets w and w . The overlap of a determinant with itself is indeed 1, as it
should, which veries the normalization of determinants. The one-electron
matrix elements also imply a spin scalar product.
N
For one-body operators F (1, 2, . . . N ) = i f (i), where f acts on one
electron, the rule is simple: determinants gives the same results as simple
product wave functions, and antisymmetry has no consequences. The expectation values are given by
|F | =
N
ui |f |ui .
(1.6)
ui
uj
vj
vi
uj
vj
vi
a)
b)
uk u uk
i
vk
c)
uj
ui
ui
d)
vk
e)
ui
uj
f)
for ui
= vi , we get the vertex c) and its exchange companion d), representing
the two terms in the expression
|F | =
N
i=k
(1.9)
This is similar to (1.8), but there is a summation over all the spinorbitals
present in both determinants, which act as background particles while the
i-th electrons jumps from v(i) to u(i). Finally, for the expectation value we
get the vertices e) and f) , that is,
|F | =
N
j=i
(1.10)
1
()P+Q
N!
P,Q
(1.11)
u1 (1)u2 (2) . . . uN (N )|f |vQ1 (P1 )vQ2 (P2 ) . . . vQN (PN ) .
Now the Q summation yields back the determinant with a permutation
P of the electrons, that is, ()P ; the ()P factor cancels the one already
present in (1.11). Hence,
2 x0
this is equivalent to
a + a
x
= ,
x0
2
ix0 p
a a
= ,
h
(1.16)
(1.17)
(1.18)
n1
0
1
2
3
n
1 + n
2
;
2
jz =
2
n
1 n
.
2
(1.19)
To extend this idea, one can observe that introducing a spinor operator
a1
=
(1.20)
a2
we may write
j=
This extends naturally to
h
;
2
jz =
h
z .
2
j =
(1.21)
(1.22)
h
(a a2 + a2 a1 );
2 1
and hence
j+ = h
a1 a2 ;
jy =
i
h
(a1 a2 + a2 a1 )
2
j = h
a2 a1 .
(1.23)
(1.24)
(1.25)
2 j(j + 1).
j2 = h
(1.26)
K = h
a 1 a2 .
(1.27)
(1.28)
U (s) = h
c
a 2
abc
b
L
2
+
c 2
L
2
a
b 2
c 2
e ( s ) +( L ) +( L ) .
(1.29)
We can calculate this exactly for large enough L and the result diverges as
0. One nds (see Appendix 1)
2
hc 2 L2
1
d
1
(1.30)
U (s) =
2
d
es 1
Using the expansion
Bn y n
1
1 y2
1 y4
y
=
1
y
+
+
...
=
ey 1
2
6 2!
30 4!
n!
n
(the Bn are called Bernoulli numbers), one obtains:
2
s
d
1
2 1
hc 2 L2
1
+
U (s) =
2
d
2
2 12s 30 4!s3
(1.31)
(1.32)
The rst two terms lead to the aforementioned divergence: should we try to
remove all the radiation from the cavity, including the high frequency modes,
that would cost us innite energy. However, the divergence disappears if we
ask: what changes if we shift one side of the cavity by 1 cm? To better answer
this question, suppose a cavity of length R is divided in two equal halves by a
mirror: evidently the energy of the vacuum is the diverging quantity 2U (R/2).
If instead the mirror is at distance s from one end and R s from the other,
the vacuum energy must be U (s) + U (R s), which also diverges. The nite
dierence
R
}
(1.33)
E(s) = lim {U (s) + U (R s) 2U
R
2
has the physical meaning of an energy that must be supplied to the system
in order to shift the mirror to the middle of the cavity. If the cavity is large,
this can be identied with the interaction energy at distance s. Eventually
one can let 0. The zero point energy decrease per unit surface is thus
hc
2
,
(1.34)
720 s3
and since the radiation pressure is proportional to the energy density one
observes an attractive force
E =
hc
2
.
240 s4
Measuring distances s in m, one nds
F =
(1.35)
0.013
dyne/cm2 .
(1.36)
s4
This force and its dependence on material and surface properties is actively
investigated and could be used to operate nano-machines.
F =
1.2.3 Fermions
The second quantization formalism for Fermions was invented in order to
deal with phenomena like neutron decay n p + e + or pair creation in
particle physics, but to create an electron-positron pair one needs about a
million eV. In condensed matter physics the typical energy scale is much
less than that, yet many important phenomena are naturally described in
terms of the creation (or annihilation) of fermion quasi-particles. Electronhole pairs can be created very much like electron-positron ones. In scattering
processes, when all the particles are conserved, one can proceed with Slater
determinants in rst quantization; however, second quantization formalism
is much easier to work with.
The change from bosons to fermions replaces permanents with determinants. In place of a N-times excited oscillator representing N bosons in a given
mode, we now consider N -fermion determinants |u1 u2 . . . uN |, where the spinorbitals are chosen from a complete orthonormal set {wi }. The index i can
be discrete or continuous but implies a xed ordering of the complete set. In
this way, one can convene e.g. that in |u1 u2 . . . uN | the indices 1 N are in
increasing order thereby avoiding multiple counting of the same state. The
zero-particles or vacuum state |vac replaces the oscillator ground state. For
the determinants, it is generally preferable
notation like
to use a compact
u
(1)
u
(2)
m
m
|um un | rather than the explicit 12 Det
which contains the
un (1) um (2)
3
same information. Consider the following correspondence between determinants and states of the Hilbert space with various numbers of electrons:
First Quantization
No electrons state(vacuum)
1 body state uk
2 body determinant |um un |
3 body determinant |um un up |
...
Second Quantization
|vac
ck |vac
cm cn |vac
cm cn cp |vac
...
(1.37)
(1.38)
3
mathematically, it is an isomorphism; it can be thought of as a change in
notation.
10
(1.39)
(1.40)
(1.41)
(1.42)
(1.43)
cm |vac = 0.
(1.44)
and
It obeys the conjugate of the anticommutation rules (1.39), namely,
Next consider
[cm , cn ]+ cm cn + cn cm = 0, c2m = 0.
(1.45)
(1.46)
np cp cp
11
(1.48)
(1.49)
(1.50)
the rule is
bn =
k
bn =
(1.51)
u
(y)u
wave function
n (x) = (x y); thus it is a perfectly localized
n n
electron. The rules are readily seen to be
[ (x), (y)]+ = 0, [ (x), (y)]+ = 0,
(1.53)
and
[ (y), (x)]+ =
p,q
up (x)up (y) = (x y)
p,q
(1.54)
where the also imposes the same spin for both spinors.
A one-body operator V (x) in second-quantized form becomes
V = dx (x)V (x) (x) =
Vp,q cp cq .
p,q
(1.55)
12
This gives the correct matrix elements between determinantal states, as one
can verify.
The above expressions imply spin sum along with the space integrals,
although this was not shown explicitly; let me write the spin components, for
one-body operators:
V =
dx V, (x)
(1.56)
,
(1.58)
which is obtained from (1.57) by taking a Fourier transform in discrete notation. A two-body operator U (x, y) becomes
(please note the order of indices carefully). The Hamiltonian for N interacting
electrons in an external potential (x) is the true many-body Hamiltonian
in the non-relativistic limit that we shall often regard as the full many-body
problem for which approximations must be sought. It may be written
(1.60)
1
h0 (i)
2i + V (ri ) =
2
i
i
(1.61)
with T the kinetic energy and Vext the external potential energy while
U=
1
uC (ri ri )
2
(1.62)
i=j
13
H = H0 + U,
H0 =
dr (r)h0 (r),
1
dxdy (x) (y)uC (x y), (y) (x). (1.63)
U =
2
,,,
Often the spin indices are understood as implicit in the integrations. It should
be kept in mind that relativistic corrections are needed in most problems with
light elements and the relativistic formulation is needed when heavy elements
are involved. Fortunately, the ideas that we shall develop lend themselves to
a direct generalization to Diracs framework.
1.2.5 Hubbard Model for the Hydrogen Molecule
The Hubbard Model is a lattice of atoms or sites that can host one electron
per spin; there is a hopping term between nearest neighbors like in a tightbinding model and a repulsion U between two electrons on the same atom.
The Hubbard Hamiltonian
cj ci + U
ni ni ,
(1.64)
H =K +W =t
i,j,
where K stands for the kinetic energy while W accounts for the on-site repulsive interaction. The summation on i, j runs over sites i and j which
are nearest neighbors in a cubic lattice. This is often called trivial Hubbard
Model to distinguish it from its extensions, involving degenerate orbitals and
o-site interactions, that have been studied for many purposes.4
To model H2 in the same spirit we represent the 1s orbitals of both atoms
= T + W
with
by two sites a and b and H
T = th
ca cb + cb ca
(1.65)
(1.66)
Some people blame the Hubbard Model and its extensions as too idealized to
be realistic. Indeed nobody would use them to rene well-understood properties of
Silicon. However, there are lots of problems involving strong correlations and e.g.
transport, spectroscopies, time-dependent perturbations, which are far too hard for
an ab-initio description. Hubbard-like models are primarily conceptual tools aimed
at a semi-quantitative understanding. We shall see particularly in Chapters 4, 5 and
10 that often they allow to deal with highly excited states of strongly interacting
system very successfully. The Bosonic Hubbard Model is also important, e.g. in the
rapidly developing subject of Cold Bosonic Atoms in Optical Lattices (see Ref. [15].
14
(1.68)
U
th
U
.
th
0110
1 0 0 1
T = th
(1.69)
1 0 0 1.
0110
One
the eigenvalues:
E = 0 for the triplet, and E0 = U, E =
nds
1
2
2
for the singlets, with E the ground state (remark16th + U
2 U
ably) for any U > 0. Magnetism never obtains in this model.
15
(1.70)
such that the interaction V takes the system to an enlarged Hilbert space,
involving extra degrees of freedom not in action in the simple problem described by H0 . Let A denote the restricted space and B the enlargement.
Typically,
HA 0
H0 =
,
(1.71)
0 HB
and
V =
0 v
v 0
(1.72)
is the mixing term. A standard way to solve such problems, that we shall
meet several times in this book, is by a canonical transformation
= U HU 1
H H
is block-diagonal:
where U is designed such that H
A 0
H
H=
B
0 H
(1.73)
(1.74)
(1.75)
(1.76)
(1.77)
(1.78)
that is,
0 v
v 0
HA 0
0 s
+[
,
] = 0.
0 HB
s 0
(1.79)
16
0 s
s 0
.
(1.80)
vn
(B)
E
(A)
En
, (s )m =
(v )m
(B)
E
(A)
Em
(1.81)
2
2 [S1 , V
(1.83)
where
2
[S1 , V ].
(1.84)
2
The eect of V can be obtained by working within the A subspace with
a renormalized Hamiltonian (see (10.48),( 1.82)) with elements
vm
vn
vn
1
vm
(Hint )mn =
+
.
(1.85)
(A)
(B)
(A)
(B)
2
En E
B Em E
Hint =
v | |v
(B)
E (A)
(1.86)
(1.87)
17
(1.88)
3 1
1 0
H =
1 0
1 0
1 0
11
00
00
00
00
1
0
0
0
nd variationally
the eigenfunctions of the form
2
2
=
. Normalizzation requires N = | = + 4 = 1 while E =
= 0 = =
18
|m|H |n |2
(0)
(0)
Em En
(1.90)
(0)
)
n = n +
(0)
(0)
(0)
(0)
Em Ek
Em Ek
k
k|H |n n|H |m
+
+
(1.91)
(0)
(0)
(0)
(0)
n (Em Ek )(Em En )
(0)
where n are the unperturbed ones. A much more general form of perturbation theory will be developed starting from Chapter 11.
(2.1)
where an extra term V (t) appears: sometimes the complication V (t) depends
on time, but in other cases it is static. Here, H(t) is in the Schrodinger picture,
i
h |S (t) = HS (t)|S (t)
(2.2)
t
which is notoriously dicult. We can solve formally by introducing the unitary time evolution operator US such that
|S (t) = US (t, t0 )|S (t0 ) .
(2.3)
20
US (t, t0 ) = 1
i
h
dt1 HS (t1 ) + (
t0
i 2
)
h
t
t1
t0
(2.5)
(2.6)
where t the domain is limited to t > t1 > . . . tn > t0 . If H does not depend on
n
n
n H (tt0 )
and the resulting exponential
time, one gets trivially In (t) = ( i
h
)
n!
series is immediately summed. This n! denominator proves useful and we can
bring it out in the time dependent case by a trick. Perform any permutation
of the time variables t1 . . . tn in In (t); in the new ordering the earlier times
will remain on the right of the later ones, while the value of the integral
remains unaltered. Thus one can sum all the n! identical replicas obtained
by permutation and divide by n!; accordingly, one denes the time ordering
operator P which puts earlier times on the right. For two terms
P [HS (t1 )HS (t2 )] = HS (t1 )HS (t2 )(t1 t2 ) + HS (t2 )HS (t1 )(t2 t1 );
with more operators,
(2.7)
Under the action of P the operators can be permuted freely as if they commuted. Actually when dealing with electron operators one uses Wicks time
ordering operator T which is dened like P except that any exchange of
fermion operators which is needed to go from the given order to the standard
earlier to the right order brings a sign. Thus, if A ad B are fermion creation
or annihilator operators, T is such that
T [A(t)B(t )] = A(t)B(t )(t t ) B(t )A(t)(t t).
When acting on Hamiltonians where Fermi operators occur in pairs P and
T have the same eect, but T permits simplifying the denition of fermion
Greens functions. Summing over the permutations, that give identical contributions, and dividing by their number, the integral (2.6) can be rewritten
t
t
1 i
In (t) = ( )n
dt1 . . .
dtn T [H(t1 ) . . . H(tn )]
(2.8)
n!
h
t0
t0
and we may formally sum the series:
US (t, t0 ) = T exp
i
h
t
t0
d HS ( ) ;
(2.9)
here T exp is a conventional notation that means nothing but the exponential
series of time-ordered products.
21
(2.10)
|H |S (t0 )
(2.11)
where by denition
is the t-independent snapshot of the golden-age , while any operator A,
including c and c , has an extra dynamic time dependence
AH (t) = US (t, t0 )AS (t)US (t, t0 ).
(2.12)
iHt
h
Ae
iHt
h
(2.13)
(2.14)
since
i
h
dU (t, t0 )
dUS (t, t0 )
= HS (t) US (t, t0 ), i
= US (t, t0 )HS (t)
h S
dt
dt
(2.15)
the result is
1
In 1 dimension, however, the Bethe Ansatz allows solving exactly some important models, see Chapter 18.
22
dAH
= US (t, t0 )HS (t) AS US (t, t0 )
dt
dAS
+US (t, t0 )AS HS (t) US (t, t0 ) + ih
.
dt H
i
h
(2.16)
= TC
exp i
t0
t0
dt H(t ) AS (t)
(2.18)
t0
t+
t
Re z
t
a)
b)
Fig. 2.1. a) A contour on the complex time z plane for obtaining A(t) from a
single Schr
odinger-picture evolution. Since along a closed path starting and ending
at t one collects US (t , t ) = 1, the path can be deformed freely as long as it starts
and ends at t0 and goes through t. b)The Keldysh contour; its main use will be
shown in Chapter (13)
23
We shall see in Chapter 10 that this is the most natural way to make
contact with experiment, and is also important for the connection with thermal physics that follows. The evolution operator satises the group propertyU (t, t1 )U (t1 , t2 ) = U (t, t2 ); hence the path can be deformed freely as
long as it starts and ends at t0 and goes through t. In time-dependent problems, the most common contour is the Keldysh one from t0 to t0 +
and back to t0 ; there are an ascending or positive branch and a descending or
negative branch, and a physical time can be taken on any of the two. I shall
write t+ and t the times taken on the ascending and descending branch
respectively; however A(t) = A(t+ ) = A(t ) .
2.2.2 Thermal Averages
The initial state of the experiment on a solid is never an eigenvector of the
Hamiltonian; it is described [118] by a Hermitean density matrix that we may
denote
wi |i i|,
(2.19)
=
i
1
KB T
eK
, K = H N,
Z
(2.20)
(2.21)
is the partition function. As detailed e.g. in [117], this yields the maximum
entropy S = KB ln() with the constraints that T r = 1, particle number
and energy must be kept xed. Thus,
(2.22)
For an independent-electron system with Hamiltonian H0 = p p cp cp ,
any energy eigenstate is specied by the set of occupation numbers np of
the one-electron levels, and the trace sums over all possible choices of np .
Each term of the sum is a product of factors e(p) from lled states
and factors 1 from empty states. Let us pick a particular level k and let X(k)
denote the contribution to Z from all the congurations with nk = 0; then, the
contribution to Z from all the congurations with nk = 1 is e(k ) X(k),
24
since the other levels give the same contribution regardless the population
of level k. We can factor (1 + e(k ) ) from the trace. Therefore we may
conclude that
(1 + e(k ) ).
(2.23)
Z=
k
t0 i
t0
dt H
] T r[eN T e
t0 i
t0
dt H
];
i
(2.24)
t0 i
dt H
A(t)
= T r0 AH (t), =
(2.26)
(2.27)
Im(t)
25
Im(t)
t0
t0
tRe(t)
t = t0 i
Re(t)
t = t0 i
a)
b)
Fig. 2.2. a) Contour for thermal averages at time t0 . The or imaginary time axis
is shown: = Im(t) is actually a real variable. b) Contour for thermal averages
at time t. Complex times in the vertical track are latest.
(2.30)
iH0 t iH0 t
AS (t)e
iH0 t iH0 t
One nds:
i
(2.32)
(2.33)
26
Today
H = I
System in
ground state
of H0
S = I
t0
satises
and
I (t) = UI (t, )I ( )
(2.34)
(2.35)
i
The solution
(2.36)
(2.37)
Problems
2.1. For H0 =
p p cp cp ,
np =
1
1
Tr(np ) =
.
Z
1 + e(p )
(2.40)
27
2.2. Let H = H0 + H1 and U0S (t, t ) the evolution operator for H0 . Write an
equation for the evolution operator US (t, t ) for H.
2.3. Let A1 , A2 denote fermion creation or annihilation operators and an
operator such as a density which commutes with A1 , A2 under Wicks T
d
ordering. Find dt
T {A1 (t1 )A2 (t2 )(t)}. How is the result generalized to several
operators A1 , An ?
2.4. Derive the useful identity [205] holding for any Hamiltonian H that
depends on a parameter
d iH (t t)
e
= i
d
t
t
d eiH (t )
dH iH ( t)
e
.
d
(2.41)
nlj EB (eV)
1s 21 7112.0
2p 12 842.0
2p 32 719.9
3s 21 706.8
3p 12 94.0
3p 32 52.7
3d 23 52.7
1
v (ri ri )
2
i=j
(3.2)
30
(3.3)
(1, 2, . . . N ) =
... ... ...
N! . . .
u1 (N ) u2 (N ) . . . uN (N )
written in terms of spin-orbitals ui (i) to be determined. This enables us to
write the average of the one-body operator according to (1.7)
|F | =
N
ui |f |ui .
N
[ui (1)uj (2)|f (1, 2)|ui (1)uj (2) uj (1)ui (2)|f (1, 2)|ui (1)uj (2) ].
j=i
i
Ii +
1
(Cij Eij ),
2
(3.4)
ij
where Ii = ui |h(i)|ui , while Cij and Eij are respectively the Coulomb and
exchange integrals involving the Coulomb interaction and orbitals i and j.
We are facing two main problems: on one hand, the determinantal wave
function is appropriate for independent electrons and on the other we need to
specify the orbitals somehow. We nd a way out of both problems if we use the
determinantal form as a variational ansatz and seek for optimal spin-orbitals;
we impose normalization by a Lagrange multiplier for each spin-orbital. If we
wish, we may use other Lagrange multipliers to enforce orthogonality, but if
we forget this requirement, orthogonal spin-orbitals are obtained anyhow.
The main reason why the HF approximation is important is that the
equations are written in the same way for all systems; we can illustrate the
procedure starting with the ground state of He. Then the two spin-orbitals
are u1s = u1s (r) and u1s = u1s (r), in obvious notation, and
E2 = u1s (1)| h(1) |u1s (1) + u1s (2)| h(2) |u1s (2)
+ u1s (1)u1s (2)| r112 |u1s (1)u1s (2)
(3.5)
does not involve exchange terms. We vary the orbital u1s (r) requiring E2 =
N, where N is the norm of u1s (r) and is the Lagrange multiplier. We
obtain the Hartree equation
h(1)u1s (1) + u1s (1)
d3 r2 u1s (2)
1
u1s (2) = u1s (1).
r12
31
(3.6)
(we speak of Hartree-Fock when exchange terms appear). Thus the optimal
orbital is obtained by a Schr
odinger-like equation where the electrons feels,
along with the nuclear potential contained in h, the Hartree potential
1
u1s (2);
(3.7)
V H (1) = d3 r2 u1s (2)
r12
V H is just the electrostatic potential due to the charge cloud of the oppositespin electron. Having chosen an independent-electron form of we are trying
to compensate the neglect of the Coulomb interaction by including its average
as an eective potential or mean eld. The Hartree approximation takes into
account the quantum nature of the electron and is self-consistent, that is, it
accounts for electrostatics. Indeed, HF is also called the self-consistent eld
method. Yet, it is still far from exact. The He ground state energy turns out
to be E2 = 77.866 eV, which is more than 1 eV too high compared to the
exact result ( 79 eV). This large discrepancy is the correlation energy and
is due exclusively to the determinantal form of . In other terms, the electron
does not see the average cloud of the other one, but a point particle with
which it can correlate its motion. The discrepancy may be small compared
to the binding energy of the system and also compared to core-level binding
energies, but since 1eV is the scale of chemical binding energy the accuracy
of the HF method is questionable if one wants to predict chemical trends.
The excited state 1s2s3 S of He can also be approximated variationally
as discussed in Sect. 1.4. Triplet He is called Orthohelium and converts to
the singlet Parahelium after a long time (spin-orbit coupling is small). The
relevant congurations are u1s u2s , with mS = 1, u1s u2s , with mS = 1.
Of course, there is also mS = 0, with the embarassing non-determinantal
conguration 12 [u1s u2s + u1s u2s ]. This state would be outside the scope
of the HF method, but we know that its energy is the same and anyhow it can
be reached from the determinantal states by a 90 degrees rotation. Hence we
can concentrate on u1s u2s , and repeating the above argument and setting
a = u1s , b = u2s we nd:
E2 = Ia + Ib + Cab Eab .
(3.8)
32
arises. However, this is a non-local potential (the r.h.s does not depend just
on the local value of a).
V ex ( x)a( x) = b( x)
b (y)a( y )
|x y|
dy
(3.10)
(3.11)
Exchanging a and b,
a| h |b = b a | b .
Taking the complex conjugate and subtracting from (3.11)we get
0 = (a b ) b | a ,
and non-degenerate orbitals are orthogonal.
The analogy of the HF equations to Schrodingers suggests that the so
called Koopmans eigenvalue a is the energy eigenvalue of the electron moving in spin-orbital a. However, one should not give any physical signicance
to the individual orbitals. One can introduce unitary linear transformations
of orbitals. Then, the determinant does not change, and nothing changes
since has a physical meaning, the spin-orbitals do not possess any by themselves. No physical observable corresponds to the energy of an orbital. Let us
take the scalar product of Equation (3.9) by |a :
a(1) b(1)a(2)b(2)
|a(1)|2 |b(2)|2
d1d2
= a ,
a(1)| h(1) |a(1) + d1d2
r12
r12
that is,
a = Ia + Cab Eab .
Using
(3.12)
33
N
1
[Cij Eij ] .
2
N
Ii +
(3.13)
i=j
(3.14)
j
(i)
where EN 1 refers to the system ionized in spin-orbital i, while the other spinorbitals remain frozen. Looking for the extremum of energy constrained by
normalization one nds the HF equations. We introduce the direct potential
d
V =
N
i
Vid (r),
Vid (r)
dr
34
V ex =
Viex (r), Viex (r)f (r) = ui (r)
dr
ui (r ) f (r )
|r r |
(3.15)
p2
Z
+ V d (r) V ex (r)
2m |r|
(3.16)
(3.17)
35
Ze2
,
|r i |
(3.18)
36
and a constant source of ideas that then are applied to the realistic calculations. The electrostatic eects are included in the theory, and they improve
it considerably. We see that what we can learn from SCF approximation.
The most fundamental issue is the cohesion of a metal piece. We consider
a large cube of Jellium of volume V containing N electrons, with N/V = n
and impose periodic boundary conditions. The method of Hartree describes
to the state of the Jellium with a product wave function of the form
(1, 2, , N ) = u1 (r1 )u2 (r2 ) uN (rN )
(3.19)
write down the direct term Vd uk (r) = uk (r) k dr |rr
| uk (r ) uk (r) and
perform the charateristic exchange
4e2
e2
ikr 1
(r )uk (r) = e
Vex uk =
u
.
(3.20)
dr
k
|r r |
V |k k|2
k
The exchange term also goes like eikr , and the Hartree-Fock equations read
[
p2
eikr
e2
4
eikr
.
=
(k)
]
2m
V |k k|2 V
V
(3.21)
(k) =
2m
(2)3
d3 k (kF k )
|k
1
k|2 .
(3.22)
Spin density wave solution of the Hartree-Fock equations are energetically favored at low n, however the energy dierence is far too small to be relevant o the
cohesion issue. Cs has low enough n to have a spin density wave in the ground state
according to Hartree-Fock equations; the actual metal, however, is not magnetic.
37
dierent from that of a free electron. The Jellium is stable if an electron at the
Fermi level is bound. This depends on the competition between the negative
contribution of the exchange term and that positive one of the kinetic energy.
The outcome of the competition depends on kF , that is an the density n of the
Jellium. For suciently small density (kF a0 << 1) the attraction prevails,
and the metal exists.
The function (k) has a logarithmic singularity for k = kF , where its
derivative diverges. This is physically wrong, no such behavior is observed,
but we have understood the reason for the existence of metals, and if we want
to understand more we must go into the many-body problem.
N
N
ei ej
2 2i
h
+
T + V.
2 i=1 mi
rij
i<j
(3.23)
The ground state wave function ({r i }) depends on the set {ri } of the N
position vectors. Consider the rescaled, normalized wave function
({ri }) =
3N
2
({r i }).
(3.24)
One checks easily that kinetic and Coulomb energies scale dierently, that is,
|T | = 2 |T | ,
(3.25)
|V | = |V | .
(3.26)
(3.27)
while
Hence,
38
(3.28)
but since the optimum value is = 1 we end up with the Virial theorem
2T + V = 0.
(3.29)
This is one of the few known exact statements about interacting many-body
problems. The Hartree-Fock ground state satises the Virial Theorem, and
an approximate self-consistent calculation can be improved by scaling.
with
H0 =
Z 2
p
Ze2
,
2m
ri
i
HC =
=
Hrel
Z
e2
,
r
i<j ij
Z
(ri )Li S i
(3.31)
(3.32)
(3.33)
(3.34)
where pi and ri are electron momenta and coordinates, rij are distances;
contains the spin-orbit coupling between orbital angular momentum Li
Hrel
and spin Si of electron i which is the most notable relativistic correction
39
(other corrections involving the orbital currents are actually larger, but less
evident since they fail to split levels). In the central eld model, the states of
the atom are constructed using 1-electron orbitals computed with a suitable
central V (r); the wave functions dier from the Hydrogen-like orbitals only in
the radial functions RnL (r). In terms of the one-electron basis states labeled
by n, L, m, mS quantum numbers2 , one starts assuming a conguration of the
atom (see Table 3.1).
H
He
Li
Be
B
C
N
O
F
Ne
Na
Mg
Al
Si
P
S
Cl
Ar
K
Ca
Sc
Ti
V
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
Cr 24 3d5 4s Ag 47 4dl0 5s
ls
2
Mn 25 3d5 4s2 Cd 48 4dl0 5s2
1s
Fe 26 3d6 4s2 In 49 5s2 5p
2s
Co 27 3d7 4s2 Sn 50 5s2 5p2
2s2
2s2 2p Ni 28 3d8 4s2 Sb 51 5s2 5p3
2s2 2p2 Cu 29 3d10 4s Te 52 5s2 5p4
2s2 2p3 Zn 30 3d10 4s2
I 53 5s2 5p5
2s2 2p4 Ga 31 4s2 4p
Xe 54
5p6
2
2
2
5
2s 2p Ge 32 4s 4p
Cs 55
6s
As 33 4s2 4p3 Ba 56
2p6
6s2
Se 34 4s2 4p4 La 57 5d6s2
3s
Br 35 4s2 4p5 Ce 58 4f 5d6s2
3s2
3s2 3p Kr 36 4p6
Pr 59 4f 3 6s2
2
2
5s
3s 3p Rb 37
Nd 60 4f 4 6s2
2
2
3
Sr 38 5s
3s 3p
Pm 61 4f 5 6s2
3s2 3p4 Y 39 4d5s2 Sm 62 4f 6 6s2
3s2 3p5 Z r 40 4d2 5s2 Eu 63 4f 7 6s2
Nb 41 4d4 5s Gd 64 4f 7 5d6s2
3p6
Mo
42 4d5 5s Tb 65 4f 9 6s2
4s
Tc 43 4d5 5s2 Dy 66 4f l0 6s2
4s 2
3d4s2 Ru 44 4d7 5s Ho 67 4f 11 6s2
3d2 4s2 Rh 45 4d8 5s
Er 68 4f 12 6s2
3d3 4s2 Pd 46 4dl0
Tm 69 4f 13 6s2
Yb
Lu
Hf
Ta
W
Re
Os
Ir
Pt
Au
Hg
Tl
Pb
Bi
Po
At
Rn
Fr
Ra
Ac
Th
Pa
U
70
71
72
73
74
75
76
77
78
79
80
81
82
83
84
85
86
87
88
89
90
91
92
4f 14 6s2
5d6s2
5d2 6s2
5d3 6s2
5d4 6s2
5d5 6s2
5d6 6s2
5d7 6s2
5d9 6s
5d10 6s
5d10 6s
6s2 6p
6s2 6p2
6s2 6p3
6s2 6p4
6s2 6p5
6p6
7s
7s2
6d7s2
6d2 7s2
5f 2 6d7s2
5f 3 6d7s2
Table 3.1. The atomic numbers and the ground state conguration of the elements.
A shell comprises all the orbitals of a given n, and the inner ones, entirely
occupied by electrons, are core shells. The inner shells contribute most of the
binding energy; their charge screens the nuclear potential and contributes in
an important way to the V (r), in which the external electrons
move. The
M
=
0, the z
magnetic quantum number of a closed shell is ML =
L
i
i
component of spin is zero and thus all angular momentum quantum numbers
2
This is a very convenient basis, but individual electron quantum numbers do not
represent observable quantities; any measurement will give the quantum numbers
of the atom.
40
vanish; the parity is +1. Thus, the parity and angular momentum quantum
numbers of the atom are determined by the outer (or valence) electrons. This
suggests that Htot can be simplied. The valence electrons are modeled by
,
H = Hv + HC + Hrel
(3.35)
where the sums are restricted to incomplete shells. The residual Coulomb
interaction HC does not contain the core electron contributions and can be
dealt with approximately as a perturbation; for light atoms Hrel
is also a
perturbation. In the limit HC = Hrel = 0 all the states in a given conguration would be degenerate; for instance, a C atom in the fundamental
2 2
2
conguration
1s 2s 2p has 2 electrons in the six p spin-orbitals and would
6
have a
= 15 times degenerate ground state. This degeneracy stems from
2
the invariance of Hv for independent rotations of electrons orbitals in space
and of the total spin. The orbital angular momentum of each electron and
the total spin are good quantum numbers in this limit. In the presence of
interactions, the only conservation laws are
[H, J ] = 0
(3.36)
J =L+S
(3.38)
41
ML + MS . Since H is invariant for independent rigid rotations in ordinary space and spin space, E is independent of ML and MS . 2) diagonalizing H, L2 , S 2 , J 2 , J z dening the |LSJMJ basis with quantum numbers
E, L, S, J, MJ . The two bases are connected by a unitary transformation, and
both schemes are referred to in literature as L-S or Russell-Saunders scheme.
The energy levels of this approximation, or atomic terms, are denoted with
symbols of the type 2S+1 L: as an example, 2 P has L=1, S=1/2; they are
degenerate (2L+1)(2S+1) times.
L-S Terms Inside a Given Conguration
The L-S terms belonging to a given conguration are resolved by HC . One
can easily nd what terms arise. For closed shells there is only 1 S. If a single
electron optical electron moves outside lled shells the entire atom has the
quantum numbers L, ML and MS = 12 of the electron. If a single electron
lacks in order to make lled shells, there is a single hole, this counts like
an electron with opposite values of ML and MS . If there are two inequivalent electrons (dierent principal quantum number n) outside lled shells,
one must sum their L and S. For instance, from two electron labeled np, n p
one builds the atomic terms 1 S,3 S,1 P,3 P,1 D,3 D. With 2 or more equivalent
electrons, the possible terms are limited by from the Pauli principle, and the
best thing is to proceed by examples.
Example: C atom (conguration 1s2 2s2 2p2 ). The closed shells may be ignored, and we must consider the conguration p2 . The 1-electron states available are (m, ) with m mL = 1, 0, 1 and = 21 ; there are 6 spin-orbitals
involved. Without the Pauli principle we would nd 1 S,3 S,1 P,3 P,1 D,3 D,
that is 1 + 3 + 3 + 9 + 5 + 15 = 36 states. Many of those terms are forbidden, since only the 2 electron determinants
(m1 1 , m2 2 ) involving dierent
6
(m, ) are allowed. There are
= 15 pairs of dierent spin-orbitals; by
2
linear combination on can form 15 allowed 2-electron states with well dened
L, S, ML , MS . The (m1 1 , m2 2 ) determinants are labelled by ML and MS ,
and those with parallel spins belong to S=1; MS = 0 is compatible with
singlet and triplet. Not all determinants have well dened L; however the determinant with the largest ML must belong to the maximum L. In this case
the largest ML is 2, and corresponds to (m1 1 , m2 2 ) = (1+, 1), having
MS = 0. Thus there is a 1 D term. 1 D has 5 states, ML = 2, 1, 0, 1, 2,
and we must nd 10 more orthogonal states, with L < 2. With ML = 1 we
can nd the 4 determinants (1, 0) having mixed L. A linear combination,
which might be found using the shift operator S on the state with ML = 2,
belongs to 1 D; 3 ML = 1 states are left, with L = 1 and MS 1; these belong
to 3 P (9 states). we have found 5+9=14 states out of 15; there is room only
for a 1 S. Thus the conguration is resolved as follows:
p2 1 S,3 P,1 D.
(3.39)
42
(3.40)
1
1
s1 m2 ms2 L1 m1 L2 m2 |LML ms1 ms2 |SMS ;
2
2
one has to normalize the result again in general (see Problem 3.3, 3.4). A
third electron can then be added if needed by multiplying by a spin-orbital
with the Clebsh-Gordan coecients, and antisymmetrizing the result. For
many electrons this build-up process becomes very time consuming, but one
can make the process faster by using fractional parentage coecients [147]
[148]. These table present the wave functions coded in a special way; this
saves labor, but introduces no new physical concepts.
3.5.2 Hunds First Rule
Hund established two empirical rules, that hold with no exception in atomic
physics, and are very popular with the students because there is no proof to
learn. This is the rst.
The lowest L-S level of the atomic conguration has the lowest S and the highest L
compatibly with S.
This is no theorem, but is true and reasonable, since high spin implies a
very antisymmetrical orbital wave function and therefore a reduced repulsion; increasing L also lowers the energy because higher L wave functions are
more diuse. For Z=6 ( C atom), the conguration is 2p2 : the terms are
1 3
S, P,1 D and the ground state is 3 P . For Z=74 (Tungsten ) the conguration is 5p6 5d4 6s2 ; the incomplete shell is 5d4 . With ML = 2, 1, 0, 1, 2, all
spins can be parallel, so the Hund rule wants S = 2. The maximum ML is
2 + 1 + 0 + (1) = 2. The ground term is 5 D. In the L-S limit (no spin-orbit
interaction) the states of a term are all degenerate. Combining the states of a
LS term with various ML and MS by means of the Clebsh-Gordan coecients
one builds |L, S, J, MJ states. The allowed values of J are obtained from L
with S by the usual angular momentum summation rule, while MJ (J, J).
Thus a term 3 F, (L=3, S=1), gives raise to 3 F4 ,3 F3 and 3 F2 , where the index to the right is J. The energy E of the atomic levels is independent of
p2
S0
D2
43
3
3
P2
P1
P0
Fig. 3.1. The p2 conguration splits into the L-S terms 1 S,1 D and 3 P , separated
by the Coulomb interaction. The relativistic correction then produces the atomic
levels as shown.
J and MJ . If we introduce Hrel
as a perturbation, the degeneracy is partly
removed; E depends on J but not on MJ . Here is the pattern of the levels
for the Carbon atom (see Fig. 3.1).
K
K
r<
YKm
(1 , 1 )YKm (2 , 2 )
1
.
= 4
K+1
r12
(2K + 1)
r
K=0 >
m=K
(3.42)
To carry out this calculation, one needs integrals involving three spherical
harmonics; these are easily obtained by codes like Mathematica or Maple;
they are tabulated in the literature [144] as cK tables, where
2K + 1 K
c (La ma , Lc mc )(m, ma mc ).
4
(3.43)
44
ab|
1
|cd = (ma + mb , mc + md )
r12
(3.44)
K=0
Actually this summation has in any case a few nonzero terms since they are
restricted by the triangular rules
|Ld Lb | K Ld + Lb
|La Lc | K La + Lc .
(3.45)
The radial integrals contain the specics of the problem, the rest is geometry.
For the diagonal elements (c,d)=(a,b) one uses the Slater integrals
F K (na La , nb Lb ) RK (a, b, a, b) (Coulomb interaction)
GK (na La , nb Lb ) RK (a, b, b, a) (Exchange interaction).
Using other widespread notations,
1
|ab =
aK (La ma , Lb mb )F K (na La , nb Lb )
ab|
r12
(3.46)
(3.47)
with
aK (La ma , Lb mb ) = cK (La ma , La ma )cK (Lb mb , Lb mb )
and
ab|
(3.48)
1
|ba =
bK (La ma , Lb mb )GK (na La , nb Lb )
r12
(3.49)
with
bK = |cK |2 .
(3.50)
HC .
The reIn this way, one can calculate the multiplet splittings due to
pulsion is invariant for space rotations and (separately) spin rotations of the
atom, which are exponentials in L, S and the HC matrix is diagonal in the
|LSML MS basis.
A closed shell contains of angular momentum L contains 2(2L + 1) electrons; the conguration with 2(2L + 1) n electrons is said to contain n
holes and to be conjugate to the one with n electrons. The expressions of
the matrices of the Coulomb interactions in terms of the Slater integrals are
the same for the conjugate congurations; however the one for n electrons
refers to the empty shell as the energy zero, and the one for n holes refers to
the lled shell. Thus,if the Slater parameters are the same in both cases, the
term separations are the same. The direct diagonalization of the Coulomb
matrix by standard methods becomes painstaking with many electrons; efcient methods based on tensor operators are reviewed by Weissbluth [147]
who also reports the results for the most common congurations. The multiplet energies for many more congurations are given by Condon and Shortley
[144] and by Slater [145].
45
2
Besides, Hrel = 2 (J L2 S 2 ) yields the rst-order correction E(J) =
(3.51)
where we have used the Wigner Eckart theorem. for example, the ne structure of the 3 P and 3 F terms of the d2 conguration depends on dierent K
constants. The Lande rule can again be derived.
The LSJM scheme is unitarily equivalent to the LSML MS scheme:
|LSJM =
|LSMLMS LSML MS |LSJM ;
(3.52)
ML MS
since the |LSMLMS vectors have the correct normalization and antisymmetry property the result is automatically normalized. For example, let us
calculate |d2 3 P0 . The relevant Clebsh-Gordan Table is the one for summing
an angular momentum 1 to angular momentum J1 ;
J
J1 + 1
J1
J1 1
m2 = 1
(J1 +m)(J1 +m+1)
(2J1 +1)(2J1 +2)
(J1 +m)(J1 m+1)
2J1 (J1 +1)
m2 = 0
(J1 m+1)(J1 +m+1)
(2J1 +1)(J1 +1)
m
J1 (J1 +1)
(J1 m)(J1 +m)
J1 (2J1 +1)
m2 = 1
(3.53)
Even in the full Dirac theory the L of the large component can be used to label
the (4-component) spinor although L fails to commute with Diracs Hamiltonian.
N
4
developing L2 = ( k Lk )2 one obtains squares of angular momenta that commute with the components and cross products that do not; similar considerations
apply to the spin operators.
46
1
|d2 3 P0 = {2|1+ , 2+ | 6|0+ , 1+ | 2|2+ , 2 |
30
(3.54)
p2
[ i + V (ri ) + Li S i ]
2m
i
(3.55)
47
J values can be mixed. In the basis of such determinants the HC + Hrel
matrix is computed and diagonalized; the eigenvectors can be labeled with J
and energy eigenvalues are MJ independent. Alternatively, one can start with
the determinantal wave functions |uL1 mL1 mS1 uL2 mL2 mS2 | and compute the
Hrel
, then perform the unitary transformation to the |LSJM basis where
HC is already diagonal (its diagonal matrix elements depend only on L and S).
We exemplify the latter procedure in the case of the d2 conguration; L = 2
for both electrons and we denote the determinants as |m1 , m2 |; for example,
|1+ , 2+ | has two spin-up electrons with m = 1 and m = 2, respectively. .
To calculate the matrix of Hrel
= [L1 S 1 + L2 S 2 ] , where is the spin
+
orbit parameter, one uses the expansion Li S i = Liz Siz + 12 {L+
i Si +Li Si };
we know how to act on the |m1 , m2 | determinants, using Equation (6.1.1).
So,
1
1
1
[Li S i ]|1+ , 2+ | = {(1 +(2) )|1+ , 2+ |+ [2|2 , 2+ |+2|1, 1|]},
2
2
2
i
(3.56)
and so on. We can calculate the eect of the spin-orbit Hamiltonian on the
|LSJMJ basis. Using the results of the problems above,
Hence,
(3.57)
(3.58)
Besides, since
1
|d2 1 S0 = [|2+ , 2 | |2 , 2+ | |1+ , 1 | + |1 , 1+ | + |0+ , 0 |],
5
(3.59)
2 1
2 3
(3.60)
d S0 |HSO |d P0 = 6.
In this way the full HSO matrix can be built; it is a block matrix since
there is no coupling of dierent J. Letting = 2, one nds the following
Hamiltonians.
For J=3 there is only 3 F3 with energy . For J=1 there is
only 3 P
1 , also
0
2
with energy . For J=4 there is the basis 1 G4 ,3 F4, and H =
. For
2 3
4 4 35 0
42 .
J=2, there is the basis 3 F2 , 1 D2 and 3 P2 and H = 4 35
0
5
42
0
1
5
48
6
1
. The degeneracy
6 0
is reduced according to the simple pattern 1 S 1 S0 , 3 P 3 P0 ,3 P1 ,3 P2 ,
1
D 1 D2 , 3 F 3 F2 ,3 F3 ,3 F4 ,1 G 1 G0 . More quantum numbers are
needed with more than 2 equivalent electrons, as discussed in Chapter 9.9.
(3.61)
a)
49
b)
Fig. 3.2. The Auger process. The primary hole is lled by either the or
the electron while the other one is emitted as the Auger electron.
hole was in the deep level ; the other was in the free particle spin-orbital
k that will be occupied by the Auger electron in the nal state. In this way,
the Auger process becomes a collision between two holes. The initial state
|i and the nal state |f of the atom are represented by 2 2 Slater
determinants; they have the same energy and are coupled by the Coulomb
interaction. The transition probability is given by the Fermi golden rule
Pif =
2
i |HC |f |2 ,
h
(3.62)
50
The Auger spectrum is plot of the current versus the kinetic energy of electrons. The Auger instruments used to measure the spectra of molecules and
solid surfaces keep the sample, the source and the electron detector-analyzer
in vacuum. The exciting source can be an electron gun or a X-ray source, like
in the ESCA (Electron Spectroscopy for Chemical Analysis) machines. When
the sample is a solid surface, in order to limit the problem of the contamination from residual gases one needs the ultra-high vacuum. The mean free
path of electrons in a solid depends on their energy, but for the transitions
that are observed commonly it is just several Angstroms; the technique is
sensitive to the surface. The peaks of the spectrum are characteristic of the
atomic species, and the Auger technique lends itself to the surface chemical
analysis using a tiny amount of material in a non-destructive way. With scanning techniques, magnied images of the surface can be obtained in which
the distribution of elements is visualized .
Actually, the atoms that belong to molecules and solids have transition
energies somewhat dierent from free atoms, and a detailed analysis of these
chemical shifts supplies further information. This can be done using CoreCore-Core features, but much more can be learned from the study of the
shapes of Core-Valence-Valence Auger lines, since the external shells are the
most sensitive to the chemical bonds. The analysis of the Auger line shapes
is currently an interesting research topic (see also Chapters 6.4,6.2 12.3. ).
The original Auger formula yields the transition energies with an error
that can be relatively small, but is typically of the order of some tens of eV.
Such an error is very large compared to the precision with which routine
measurements can be done in a modern apparatus. The main problem is that
in the nal state there are two holes, and is necessary to account for their
repulsion; the hole-hole interaction shifts the peaks of the spectrum to lower
kinetic energy and splits them into multiplets. Moreover, every diagrammatic
transition (that is, a line which is predicted from the previous arguments)
has in reality various satellites, that correspond to excites states of the nal
ion. The Wentzel theory is based on a number of assumptions that limit its
validity; it neglects the correlation eects and Relativity. The detailed theory
of the Auger spectra is necessarily involved, even in the case of free atoms.
Nevertheless, the Auger spectroscopy is one of the most important and direct
methods for the study of the correlation eects in molecules and solids.
3.7.1 Auger Selection Rules and Line Intensities
Selection rules arise in the two-step model from the conservation of J 2 , Jz
and parity between the initial, core-hole state |i and the nal state |f
including the Auger electron. In the L-S approximation, L2 , Lz and S 2 , Sz are
also conserved, while in the jj scheme, the states are labeled by the j quantum
M5 N4,5 N4,5
M4 N4,5 N4,5
3
G4 + D2
3
3
S0
51
F4 +1 S0
1
F2,3
P0,1
G4 + D2
P2
P0,1
360
Kinetic energy (eV)
365
370
Fig. 3.3. Sketch of the M5 N4,5 N4,5 and M4 N4,5 N4,5 spectrum of Cd vapor, in
arbitrary units, from measurements by H. Aksela and S. Aksela (Ref. [161]). Many
of the multiplet terms are well resolved. The assignements were done by intermrdiate
coupling calculations of line positions and intensities.
If there are open valence shells, each two-hole nal state becomes a multiplet
when valence and core-hole angular momenta are recoupled; to a rst approximation
we can often neglect such complications since usually the splitting is small.
52
In the non-relativistic limit the Auger intensities are computed by evaluating the matrix elements of HC with Pauli spinors; as a simple approximation,
Coulomb waves are used for the Auger electron states (since the Auger electron leaves a doubly charged ion behind) and Hartree-Fock orbitals for the
discrete levels. However, conguration mixing is important. For instance, in
computing KLL spectra, the mixing of 2s2 2p4 1 S0 and 2s0 2p6 1 S0 has an
important eect on the wave functions and hence on the intensities. The calculation simplies if one can approximate |i and |f as two-hole states,
and this is often useful (for instance, in noble gas and in transition metal
spectra). One then often uses as a further simplication the mixed coupling
scheme which consists in treating |i in the jj coupling scheme but |f in
the LS one. This is useful for core-valence-valence spectra in a wide range of
Z, when the spin-orbit interaction is small for valence holes but is important
for deep states. For quantitative work, we need (particularly for intermediate
and high Z) the relativistic theory with the Breit interaction
WB (1, 2) =
e2
(3.63)
where is Diracs velocity and h is the energy dierence between the scattering states (namely, the deep hole and Auger electron states). Codes are
now available; GRASP[158] (General-purpose Relativistic Atomic Structure
Program) computes Dirac-Fock orbitals, takes linear combinations of determinants with the correct J 2 .Jz , parity and seniority number labels; further
it takes linear combinations of such states, thus doing a partial conguration
mixing, and includes corrections like the eects of nuclear size and the main
QED corrections.
Problems
3.1. Find how the ground conguration of N is resolved into L-S terms.
3.2. Find how the ground conguration of Ti is resolved into L-S terms.
3.3. For the conguration d2 obtain |3 P ML = 0, MS = 0 as a combination
of determinants |m
L1 , mL2 | of one-electron wave functions, using the ClebshGordan coecients.
Note that using JMJ 1|J |JMJ = h
J(J + 1) MJ (MJ 1) one nds
for d states
L+ |2 = 0
L |2 =2|1
+
L |1 =2|2
L |1 = 6|0
+
L |0 =
6|1 L |0 = 6| 1
+
L | 1 = 6|0 L | 1 = 2| 2
L+ | 2 = 2| 1 L | 2 = 0
(3.64)
53
3.4. For the conguration d2 show how to obtain |3 P ML , MS for the other
values of ML , MS using the results of the previous problem, the shift operators
and (6.1.1). Write down all the 9 states.
3.5. Using the results of the previous problem for |3 P ML = 0, MS = 1 in
the conguration d2 verify explicitly that it is really 3 P with the specied
quantum numbers.
n dS
2 d3 x =
V
(4.1)
(where
n is the outgoing normal to the surface S bounding the volume V )
is obtained from the divergence theorem
div A d3 x =
A
n dS,
(4.2)
V
r ) )(
r)=0
(4.3)
( 2
+ V (
2 r
dened in some volume V with some boundary conditions; it is often convenient to change it into an integral equation by a method which is familiar
from classical physics. One introduces a Greens function satisfying
1
( 2
r ) = (
r
r ),
(4.4)
r ) )G(
r ,
+ V (
2 r
multiplies (4.3) by G(
r ,
r ) , (4.4) by (
r ) and subtracts; the result is
(exchanging r with r )
1
2
r , r )2
r G(
d3
(
r
)
(
r
)
G(
r
,
r ) . (4.5)
(
r)=
r
r
2 V
The V integral can be changed to a surface integral by Equation (4.1).
56
= 2
h
i|H (Ei + H)H |i ().
Transforming from frequency to time,
d it
(t) =
e
() = i|H ei(HEi )t H |i
2
= i|eiHt H eiHt H |i = i|HH
(t)HH
(0)|i .
(4.7)
(4.8)
(4.9)
we shall also need more complicated operator averages, e.g. in Sect. 6.4.1 .
57
(4.10)
Along with ph (t) = ap(t)aq (t)ar(0)as (0) and hh (t) = ap(t)aq(t)ar (0)as (0)
we shall need more complicated objects; however, simpler objects are also extremely useful. Let us start with the one-body ones. Such are the so called
lesser and greater Greens functions (ground state averages)
#
<
(t, t ) = i (t ) j (t)
(4.11)
gi,j
$
>
gi,j
(t, t ) = j (t)i (t )
(4.12)
One is obtained from the other by exchanging the operators, and in equilibrium g < informs us about the lled states, while g > knows about the empty
ones. From these, we can build retarded and advanced Greens functions; now
what matters is the order of times:
<
>
igri,j (t, t ) = (gi,j
(t, t ) + gi,j
(t, t ))(t t )
while
(r)
>
(4.13)
(4.14)
(a)
The two are related by gi,j (t, t ) = gj,i (t , t). How do they depend on
|0 ? In no way! Averaging on the vacuum, g < = 0 and
$
#
r
i gi,j
(t, t ) = vac |cj (t)ci (t ) |vac (t t )
(4.15)
and since H acting on the vacuum yields 0, we obtain
$
#
r
i gi,j
(t, t ) = vac |cj eiH(tt ) ci |vac (t t ) ,
(4.16)
(4.17)
58
This
is a one-body
problem, that can be rewritten in rst quantization
#
$
i
j| zH+i |i . Averaging on a completely lled system, g > = 0 and creation and annihilation operators are interchanged, but the result is the same
one-hole amplitude. For a partially lled system, g < brings information on
the occupied states and g > on the empty ones, but the end result is again
i
a one-body matrix element of zH+i
between spin-orbitals irrespective of
(r) (a)
can always be computed with the
where the Fermi level is. Thus, g , g
average performed over the vacuum, and are a one-particle property. Both
are characterized by the fact that they do not know where is the Fermi level
(for non-interacting systems at least; in the presence of interactions they may
have some smell of a change of occupation numbers through the change in
potential it introduces).
g (r) , g (a) are not suited for perturbation theory; the diagrammatic method
(Chapter 11 ) works with the time-ordered ones
(T )
(4.18)
(4.19)
This discontinuity is needed; we shall see (Chapter 10) that it implies a source
term in their equation of motion. In eld theory every scattering event is
represented as the annihilation of the ingoing particle and the creation of
the outgoing one: here is where the source is necessary. In real space the
time-ordered Greens function is
(T )
(4.20)
with the operators in the Heisenberg picture (the spin indices are often omitted when they are not needed).
4.2.3 Quantum Averages
The average of one-body densities f(x) = f (x) (x) (x) is
(T )
f = i lim+ lim
f (x)g (xt, x t ).
(4.21)
t t
x x
(
x , t) = i lim
lim
t t+0
x
x
(T )
(xt, x t );
g
(4.22)
[px (x x ) + (x x )px ]
j(x)
=
2m
59
(4.23)
lim
x , t) =
lim (
J (
x
x )g (xt, x t ).
2m t t+0
x
x
(4.24)
The average of one-body operators A = dxa (x) (x) (x) is
(T )
= i lim lim
A
a (x)g (xt, x t ),
(4.25)
dx
t t+ x x
upper sign for Bosons. This may be thought of as a trace over space and spin
variables:
= i lim lim T r a (x)g (T ) (xt, x t ) .
A
(4.26)
t t+ x x
Finite T rule: A
(4.27)
(T )
g (xt, x t ) =
i
T rT [ (x, t) (x , t )], Z = T r;
Z
(4.28)
(4.29)
60
=
A
dx
lim
x x, +
(4.30)
To see that, one notes that,by the cyclic property of the trace,
dxa (x)Tr[eK eK (x)(x)];
Tr[aeK (x)(x)eK ] =
K
eK commutes with = e Z and
dxa (x)Tr[ (x)(x)] =
In this way the common value of and disappears from the calT rA.
culation of the observables and the behavior of G for far from has no
physical meaning. G (T ) (x, x ) is periodic:
G (T ) (x, x ) = G (T ) (x, x + ), >
G (T ) (x, x ) = G (T ) (x + , x ), <
(4.31)
(upper sign for bosons). Indeed, let > ; then Z G (T ) (x, x ) is given
by T r{eK (x, ) (x , )}; using the cyclic property of trace this may
be transformed to read
T r{ (x , )eK (x, )} = T r{eK e+K (x , )eK (x, )}.
Thus, we got (x , + ), standing on the left of , in agreement with the
fact that + > , but this implies a sign change for Fermions, a - sign
comes from the denition of G (T ) and the rst line results; the second follows
in a similar way.
G (T ) is useful only in time-independent problems; then it is actually
G (T ) (x, x , ). According to the above discussion it is a function of
one variable dened in (, ) which can be extended to the neighboring
intervals like a periodic (antiperiodic) function for Bose (Fermi) systems. For
all practical considerations, we may extend G (T ) in this way to the real axis,
since this has no physical implications but allows to write the Fourier series
G
(T )
1 (T )
(x, x , ) =
G (x, x , n )ein t ,
n=
(4.32)
1
2
n
,
d ein G (x, x , ).
(4.33)
(4.34)
61
r+
V
( r ), such that H0 k (x) = k k (x); in second-quantized form, H0 =
= 1, the annihilation operator
k k nk ; in the Heisenberg picture, setting h
for spin-orbital a evolves with
ca (t) = eiH0 t ca eiH0 t .
From the equation of motion ca = i[H0 , ca ] , since [nk , ck ] = kk ck , one
obtains for energy eigenstates ck = ik ck and so
ck (t) = ck eik t .
(4.35)
(4.36)
where the average is taken over the ground state. For an empty (totally
full) system, this reduces to the retarded (advanced) Greens function, but
otherwise ga,b propagates both electrons and holes; for t1 < t2 an electron
is added at t1 , propagates forwards in time and is annihilated at t2 , for
t1 > t
2 a hole is introduced, propagates backwards and is annihilated. Using
bn = ak < ak |bn > (Equation 1.51) one readily nds that
k
ga,b (t) =
(4.37)
(4.38)
where nk is the occupation number. The propagator evolves with the same
phase factor as the wave functions but has a discontinuity at t = 0 and
satises the equation of motion
0
ei(+x)t dt =
i
+x+i ,
0
ei(+x)t dt =
i
+xi ,
+0,
(4.40)
62
k,k
,
k + ik
(4.41)
with k = + for empty states and k = for lled ones, and stands
for a positive innitesimal. Thus one has an electron for k > 0 and a hole
otherwise. In the space representation the retarded Greens function (4.17)
reads
k (
1
r )k (
r )
=
r|
|
r
r ,
r , ) =
(4.42)
g (r) (
k + i
H + i
k
It obeys the same equation of motion as the time-ordered one, the dierence
arises from the boundary conditions. Indeed, setting z = + i, using
1
(z H)|
r
zH
r
r ) = z
r|
=
r |
r = (
r|
1
1
|
r
H|
r (4.43)
r|
zH
zH
and exchanging
r with
r one nds
1
r ))g (r) (
r ,
r , ) = (
r
r )
( ( 2
+ V (
2 r
(4.44)
(4.45)
Using ck (t) = eik t with the substitution it and introducing the chemical potential one gets: ck ( ) = e(k ) ck , ck ( ) = e(k ) ck and so
(4.47)
where n(k) is the Fermi distribution (2.40). This is similar to (4.38); the
changes are: i becomes -1 (a mere convention), energies are referenced to the
chemical potential, it at the exponent, the energy step functions are
smoothed according to the Fermi distribution and the Wick T now acts on
the vertical track. Next, we nd the thermal Greens function in frequency
space by (4.34) and (2.40)
G(k, n ) =
1
,
in + k
(4.48)
63
Non-relativistic Bosons
Consider a Bose eld which obeys the harmonic wave equation
2
k (t) = k2 k (t) .
t2
(4.49)
2
2
2 2
2 2
c (r, t) = 0.
[ 2 +c k ]k (t) = 0
t
t2
(4.50)
ak (t) + ak (t)
ak exp [ik t] + ak exp [ik t]
=h
2k
2k
(4.51)
(4.52)
2 + 2 2
k
k k
k
1
hk ak ak +
.
2
(4.53)
(4.54)
Dk (t) =
#
$
#
$
(t) 0|ak exp [ik t] ak |0 + (t) 0|ak ak exp [ik t]|0
2k
. (4.55)
Therefore,
Dk (t) =
1
exp[ik | t | ]
{(t)exp[ik t]+(t)exp[ik t]} =
.
2k
2k
(4.56)
1
t Dk (t) = 2k {ik (t)exp[ik t] +
= i
2 {(t)exp[ik t](t)exp[ik t]}
ik (t)exp[ik t]}
(4.57)
64
is not, and
ik t
i t
2
i
ik (t)e k } 2i
t2 Dk (t) = 2 { (ik ) (t)e
ik t
i t
k
=
+ (t)e k } i (t)
2 {(t)e
{ (t) + (t)}
that is,
2
Dk (t) +k2 Dk (t) = i (t)
(4.58)
t2
Therefore,the propagator is Greens function of the wave equation. Fourier
transforming, one nds
[ 2 +k2 ]Dk () = i.
(4.59)
However the solution is problematic, because of poles on the real axis, and
the transform of (4.56)
0
1
dt ei(k )t + dt ei(+k )t
dt Dk (t) ei t =
Dk () =
2k
0
1
Dk () =
dt ei(k +i )t + dt ei(+k i )t
2k
0
1
1
1
+
=
2k i ( k + i) i ( + k i)
1
1
i
+
(4.60)
=
2k ( k + i) ( + k i)
and nally
Dkk () =
ikk
.
2 +2k i
(4.61)
For phonons and other non-relativistic bosons commonly one denes, (understanding the vacuum average)
#
$
k (t) = i P k (t) k (0)
D
(4.62)
in terms of
1
2k
(4.63)
in (4.51); then
(4.64)
65
mc 2
2
)
+
(
=0
c2 t2
h
(4.65)
i
2 + (chk)2
+ m2 c4 i
(4.67)
p =
h(k, i )
c
D (p) =
(4.68)
i
2
(cp) + m2 c4 i
(4.69)
p2
i
.
i
(4.70)
(4.71)
1
= 0,
c t
(4.72)
(4.73)
(4.74)
66
D, (x, x ) = i,
d4 p 4 eip(xx )
)
;
(2 p2 i0
(4.75)
this is evidently a Greens function of the wave equation. Actually the Greens
function is not unique and the gauge invariance allows to add to the r.h.s. of
2
(4.75) xx f (x x ), where f is an arbitrary function. The gauge in which
(4.75) holds as it stands is called the Feynman gauge.
4.3.1 Greens Functions for Tight-binding Hamiltonians
The tight-binding model Hamiltonian
ci cj , th > 0
H = th
(4.76)
<i,j>
of i. Here I consider regular linear, square and cubic lattices. Using continuum normalization, and setting the lattice parameter to 1, the Bloch energy
eigenfunctions are
1 iki
|k =
e |i .
(4.77)
2 i
In d dimensions, the energy egenvalues are:
k = 2th
d
cos k .
(4.78)
d () =
k
|0 | k | ( k ) =
1
(2)d
d
d k
BZ
2th
d
,
cos (k ) .
(4.79)
where the integral extends to the Brillouin Zone. Converting the ( k )function into a sum of momentum functions, a factor |k k | appears in
the
To
better understand the physical meaning, dierentiate
denominator.
(
p + k )2 + V (
x ) u
x ) = k u
x ). (Equation 8.23:)
(
(
2m
k
k
(
p + k )2
(
p + k )2
+
+ V ( x ) (k )u
u
x)
( x )(k )
(
k
k
2m
2m
x ) + k k u
x ). (4.80)
= (k )k u
(
(
k
k
67
Multiplying by u
| and recalling (8.22) one obtains
k
p
1
x )| |
x ) = k k .
(
(
k ,
m k ,
h
(4.81)
1 Dimension
The band edges = 2th are the extrema of k = 2th cos k at k = and
k = 0 and = 0 if || > 2th . The argument of the in (4.79) vanishes for
);
k = k = arccos( 2h
1 () =
dk
0
1
(k k )
=
.
2th sin(k)
2th sin k
(4t2h 2 )
,
4t2h 2
(4.82)
(4.83)
which in fact when integrated over all yields 1. The band-edge divergence
1 ()
0.1
0.05
-2
Fig. 4.1. Density of states on the 1d tight-binding lattice. Note the Van Hove
singularities at the band edges .
is characteristic: it reects the fact that at the band edges the group veloc
vanishes; this, combined with the one-dimensionality, leads to
ity vg = k
the inverse-square-root singularity. The symmetry around the band centre is
typical of bipartite graphs. These are lattices with the property that all
sites can be painted red or blue in such a way that any red (blue) site has
only blue (red) rst neighbors; in the linear chain these are odd and evennumbered sites. Changing the sign to all the blue orbitals is just a gauge
transformation and cannot change any physical quantity, yet it is equivalent
to sending the o-diagonal one-electron matrix element th to th . However,
k is proportional to th and must change sign as well. This can happen in
68
just one way: the spectrum is symmetric and the eigenfunctions at k and
k get exchanged by the gauge transformation.
The o-diagonal elements gm,n of the resolvent may be found by taking
1
1
= 1+ H
H between site 0 and n
= 0.
matrix elements of the identity H
One nds
q + 1 = 0.
th
Some care is needed to choose between the roots
q2
q =
where
2th
2
) 1
2th
(4.85)
(4.86)
(4.87)
( 2th )2 1 is imaginary for in the spectrum between 2th and 2th .
a branch cut on the real axis ; taking z = |z| exp[ 2i arg(z)] with the cut
along the negative Re(z) axis,
i |( )2 1| above the axis
z 2
2th
(
) 1=
i |( )2 1|
2th
below
2th
( 2th ) 1 > 0, > 2th becomes negative for < 2th and so |q | 1
everywhere. Therefore, everywhere,
g0,n () = g0,0 ()q ()|n| .
(4.88)
2 Dimensions
Using Equations (4.79,4.82), one nds
2 () =
1
(2)2
dkx
that is
1
2 () =
2
(4.89)
dk1 ( 2th cos (k)).
(4.90)
69
This integral can be written in terms of the complete elliptic integral [44]
of the rst kind. The band now extends from 4th to 4th , and is symmetric
(the graph is bipartite). Singularities become milder when integrated over,
however they are still evident (see Figure ??). The worst is the logarithmic
1
for k 0.
Van Hove singularity ( at = 0 the integrand of (4.90) goes like |k|
In other terms, this is just the singularity of 1 in integrated form, and is a
universal feature of d=2 lattices.) In addition, 2 jumps discontinuously to
2 ()
0.1
0.05
-4
-2
Fig. 4.2. Density of states on the 2d tight-binding square lattice. Note the Van
Hove singularities: jumps at the band edges and a diverging cusp a the centre.
0 at the band edges, which represents a milder singularity (in general, this
term involves a point where the function is not analytic, not necessarily a
divergence). This can again be read o (4.90): 2 = 0, for > 4, but setting
there is an
= (4 )th the band-edge singularity of 1 enters again since
interval near k = 0 where cos k > 1 /2; this interval is of order and the
integrand there is of order 1 , so the result is nonzero up to the edge.
3 ()
0.1
0.05
-6
-4
-2
70
3 Dimensions
The simple cubic lattice is bipartite, hence the density of states is again
symmetric around = 0. We can nd it by calculating
1
3 () =
2
dk2 ( 2th cos (k)).
(4.91)
(4.92)
G( r , r ) = G0 ( r , r ) d3
r G0 (
r ,
r )V (
r )G(
r ,
r ) (4.93)
where V is the perturbing potential, which is nothing but (4.92) and a special
case of the Dyson equation (see Section 11.4 below).
4.3.3 t matrix
Another form of (4.92)
1
1
1
1
=
H1
+
zH
z H0
zH
z H0
(4.94)
(4.95)
1
t.
z H0
(4.96)
71
II
Fig. 4.4. Left:Embedding of region I into II; Right: a pill-box element V between
S and its slightly inated, dotted version.
(
r)
r I
( r )
r II
( r ) =
(
r ) = (
r)
r S;
(4.97)
|H|
|
(4.98)
stationary, with
1
H = 2 + V (r).
(4.99)
2
We write in II as a functional of and reformulate the problem such that
the latter is the only unknown. Thus, the solution in I yields the solution
everywhere. Let us see how this is done in practice.
72
|H| =
d r (
r )H(
r)+
3
d r (
r )(
r)+
3
II
dS
(4.100)
2 =
2 = (
n outer dS +
n inner dS),
V
1
dS =
2
S
(
r ) ( )
n dS.
(4.101)
d3
r )(
r ).
(4.102)
r (
| = d r ( r )( r ) +
I
II
( 2 )
( ) = ||2 .
2
2
Next, integrating over II and using Greens theorem one nds
n .dS
||2 d3 r =
2
II
S
(4.103)
Here a - sign comes from the outgoing normal from I which is the opposite
of the outgoing normal from II. To simplify matters, we vary in such a way
that over S; then
0 in S and we dispose the rst term on the
73
r ))g(
r ,
( ( 2
r , ) = (
r
r )
(4.104)
+ V (
2 r
which is the negative of the retarded Greens function satisfying (4.44).Therefore,
he nds for
r II
g(
r , r )
(
r)=
r ( r )
2 S
2
(
r )
(4.105)
r g( r , r ) n d r
( r ) =
g(
r ,
r ) (
r )
n dS
(4.106)
2 S
then put
r on S and invert the relation to read
n = 2 dS g 1 (
r ,
r )(
r ),
r ,
rS
(4.107)
(
r )
S
(4.108)
74
(t),
H (t) = AF
(4.109)
d
(t) = [H(t), (t)] .
dt
(4.112)
d
eq ].
(t) = [HS , (t)] + F (t)[A,
dt
(4.113)
The solution is
i
h(t) =
e h
HS (tt )
(4.114)
(4.115)
Now we move the left exponential to the right (by the cyclic property of the
t ) in the Heisenberg picture (with HS ):
trace) thereby obtaining B(t
t
1
eq ]B(t
t )F (t )dt
Tr
B(t) =
[A,
(4.116)
i
h
i
hBA (t) = T r{Aeq B(t)
(4.117)
using again the cyclic property, we can change this into a more elegant equilibrium thermal average, namely the Kubo formula[29]
A eq AB(t)}
i
hBA (t) = T r{eq B(t)
= [B(t),
A(0)] .
75
(4.118)
0
h )
dA(i
d,
dt
(4.119)
eHS ] B(t)},
i
hBA (t) = T r{[A(0),
one nds the further Kubo formula
dA(i
h )
BA (t) =
B(t) d.
dt
0
(4.120)
This formalism has many applications; for example, if the system is exposed
to an electric eld, A is proportional to the position vector components of the
is the curelectron and the time derivative is proportional to the current, B
rent and the conductivity tensor. For the frequency-dependent conductivity,
H and using the cyclic property again, and
(4.123)
where W0 is the ground state energy of H0 . Note that unlike all the other
Greens functions, here the average is taken on the non-interacting ground
state. R is related to the interacting ground state energy E0 through the
following device. Inserting a complete set of H eigenstates,
2
| |n | ei(En W0 )t ,
(4.124)
R (t) =
n
76
we can isolate the ground state energy from (4.124) by giving the time t an
imaginary part: t t i, > 0. In this way, the exponent has a real part
t (W0 En )
which is largest for n = 0. In the long run, the ground state dominates:
R (t) = | |0 |2 ei(E0 W0 )t ,
t (1 + i) .
d
log(R(t)).
t(1i) dt
lim
(4.125)
Since UI admits the T exp expansion (2.37), the methods of Chapter 11 will
provide a practical way to compute R.
(4.126)
emphasizing that it is the average over the N-body interacting ground state
of the Hamiltonian H and that the operators are in the Heisenberg picture. For time-independent H, setting h
= 1, (x, t) = eiHt (x)eiHt , and
77
(4.127)
and the eigenstates with one more particle come into play; for t < t ,
iG(x, x , t t ) =
ei(EN 1,n EN,0 )(tt )
n
(4.128)
(4.131)
(4.132)
The real and imaginary parts of Greens functions are Hilbert transform pairs.
78
(4.133)
i Kn
(e
n|(x)|m m| (x )|n eimn (tt )
Z mn
(4.134)
(4.135)
exchanging m with n in the last line and Fourier transforming one readily
arrives at
g (r) (x, x , ) =
1 Rmn (x, x ) Kn
(e
+ seKm );
Z mn + mn + i
(4.136)
1 Rmn (x, x ) Kn
(e
+ seKm ).
Z mn + mn i
(4.137)
s eKm
1
eKn
).(4.138)
Rmn (x, x )(
Z mn
+ mn + i + mn i
Equations (4.130,4.131) can be extended to nite T, introducing the Tdependent spectral function
(x, x , ) =
2 Kn
e
Rmn (x, x )( + mn )(1 + se ) (4.139)
Z m,n
(4.140)
(4.141)
G(x, x , ) =
d ns ( )(x, x , )
[
2 ( + i))
(x, x , )
ns ( )
s
]
( i)
79
(4.142)
with n
() 1 ns () and ns () = (1 + se )1 The knowledge of (also
a function of T ) determines all the one-body Greens functions. For independent electrons, g (r) and (4.139) are purely one-body properties independent
of the Fermi level; therefore the temperature dependence of (4.139) is neglected usually in Density Functional and other self-consistent calculations.
This simplication may be wrong when dealing with strongly correlated systems.
4.6.3 Fluctuation-Dissipation Theorem
The above results imply a number of relationships involving and the real
and imaginary parts of the various Greens functions. One shows (Problem
4.2)the following. Assuming that Rmn (x, x ) is real (it must be positive for
x = x and it can be taken real anyhow in the absence of magnetic elds)
show that
s
d ImG(x, x , )
ReG(x, x , ) =
(tanh
) , s = . (4.143)
2
Historically such relations are known as Fluctuation-dissipation theorem.
Problems
4.1. Prove the identity
eH ] = eH
[A,
0
e H [H, A] e H d,
(4.144)
82
kC,
while
Hh =
{Vk ak a0 + h.c.}
(5.2)
k,
k,
and would represent 1 electron instead of representing any number of noninteracting elements. The two problems are closely related, and, for pedagogical reasons, I rst present the elementary one-body term and then the
many-body equation of motion approach (this is actually a one-body problem
but the equation of motion method lends itself to the many-body case.)
5.1.3 One-body Treatment
In the elementary approach one writes:
H |0 = 0 |0 + V0k |k
k
H |k = k |k + Vk0 |0 .
The H eigenstates are expandable on the old basis:
| = |0 0 | +
|k k | .
k
(5.4)
(5.5)
83
(5.6)
(5.7)
n ;
(5.8)
the basis change (5.5) entails new creation operators for states such that
a = a0 0| + k ak k|
a = a0 |0 + k ak |k
with the inverse
(5.9)
ak = < k|> a
a0 = < 0|> a
[a , n ]/
=(, )[a
, n0 ] = (, )a a a =
(, )
1 a a a =(, )a ,
da
= a , H
= a
dt
a (t) = a (0) ei t
(5.10)
<|k > ak , H] =
(<|0 > a0 + <|k > ak )
k
(5.11)
84
Now,
a 0 , a0 a k = a 0 a 0 a k a 0 a k a 0 = a 0 a 0 a k + a 0 a 0 a k = a k
Vk0 ak ,
[ak , H] = k ak + V0k a0 .
(5.12)
Vk0
|0 ;
k
(5.13)
substituting into the rst one gets an algebraic equation for the discrete
eigenvalues
0
|Vk0 |2
= 0.
k
(5.14)
85
ctitious box. This is not the way to the continuum. A brute force approach
fails because of the irrelevant complications it produces.
0 -
|Vk0 |2
k k
0
Fig. 5.1. Graphical solution of Equation (5.25). The true continuum C, before
placing the system in the box, extends from = 0 to = 6; a dot marks a root
below it. Here the continuum is replaced by equally spaced roots.
(5.15)
E
Hence, E12 d() is the number of states in the interval E1 < < E2 ;
it may diverge in the continuum limit, when N ,
but we can convert
3
86
The local (or projected) density of states (LDOS) measures the degree of
mixing of the eigenstates at energy with the 0 orbital,
( )|0| |2 .
(5.16)
0 () =
(0)
0 ()
(5.17)
where stands for a positive innitesimal, or if you like = +0 (intelligenti pauca). This is the Fourier transform of causal1 operator G(t) =
eiHt (t), and is analytical in the upper half plane.
Note that
G00 () 0|G()|0 =
|0| |2
,
+ i
(5.18)
and so
1
0 () = ImG00 ().
(5.19)
Here it is understood that C d + D , where D the possible
discrete eigenvalues outside C. If |0
= 0, a pole exists just below the real
2
axis at = i, and the residue is ||0 | . In second quantization one
denes the Greens function as the vacuum average
G00 = vac|a0
1
a |vac .
H + i 0
(5.20)
Inserting a0 = a |0 , one nds (5.18) again. However the second quantized formulation lends itself to the many-body treatment when interactions
are included.
5.1.5 The Resolvent
= 1, that is,
We obtain all the elements of G from the identity ( H + i)G
|0 = 0 | 0 = 1
0| ( H + i)G
|0 = k | 0 = 0
k| ( H + i)G
|k = 0 | k = 0
0| ( H + i)G
|k = k | k = (k k ) .
k| ( H + i)G
1
87
(5.21)
1
,
0 ()
k
(5.22)
|V0k |2
k + i
(5.23)
is the self-energy. For the other elements of the resolvent matrix see Problem
5.3.
Unlike the sum in (5.25) this is a smooth complex function of ,
k
stands for an integral;
1 () Re() = P
2 () Im() =
|Vk0 |2
k
k
|Vk0 | ( k ).
(5.24)
88
and write
ImG00 =
d ( ) |0| |2 ( ) = ()|0| |2 ,
(5.27)
Re z
cut
Resonant
state
Localized
state
Fig. 5.2. The singularities of G00 (z) must be in the lower half of the complex z
plane. Outside the continua one can nd poles close to the real axis at localized
states. Branch cuts correspond to the continua; a and b are branch points; poles
below the cuts yield resonant states.
|0| |2 =
0 ()
ImG00 ()
=
,
()
()
(5.28)
89
where the last equality holds in the thermodynamic limit. Now we can solve
(5.7) in the sense of distributions. The most general real solution of xf (x) = C
is
1
f (x) = C P + Z(x) ,
x
whereZ is a constant. Thus we nd a family of solutions:
1
|k = |0 Vk0 P
+ Z( )(k )
k
(5.30)
0 1 ()
,
2 ()
(5.31)
1
1
+
HAB GBA .
HAA
HAA
(5.33)
(5.34)
1
1
HAA HAB H
HBA .
BB
(5.35)
GBA =
Hence,
GAA =
Thus we can work within subspace A provided that we work with an energydependent eective hamiltonian
Hef f () = HAA + AA , AA = HAB
1
HBA .
HBB
(5.36)
90
(5.37)
d band
-A
Fermi
level
-I
Fig. 5.3. Relative position of energy levels in Newns model (not to scale). is the
work function. The bending of the Vacuum level is due to the surface electric eld.
1
c0 | ,
H
91
(5.38)
|0 | |
;
+ i
(5.39)
this is already familiar except for the appearance of the Fermi function f .
For the empty states, one denes the one-electron Greens function
1
(5.40)
c0 | .
H
From the Fano model we can obtain interesting quantities like the level population
n0 = 2
EF
d0 ()
(5.41)
where EF is the Fermi energy and the factor 2 is due to the spin.
Interacting Anderson Model in Mean Field Approximation
For the Hydrogen atom, the ionization potential I = 13.6eV and the electron
anity A = 0.7eV, are so dierent that it is impossible to set up a sensible
2
model without the interaction U 0 0 | re12 |0 0 |I A|. A work
function (Typically 4.5 eV) separates the Fermi level from the Vacuum
Level (minimum energy of a free electron far from the metal); so even is
smaller than I A. We need a simple calculation to orient ourselves. In the
Hartree-Fock approximation, one assumes a ground state of the determinantal
form
| =
<EF , <EF c+ c |vac .
Averaging the one-body operator
n0 =
c+
c | 0 >< 0| >
(5.42)
(5.43)
Moreover, since the anti-commutation relations allow to lump all the up spin
operators on one side, it is readily seen that
|n0 n0 | = |n0 | |n0 | n0 n0 .
(5.44)
92
Thus,
E = |H| =
(5.45)
n0
=
|0 0 | = |0 0 | .
(5.46)
(0)
(0)
0 0 + U n0 .
(5.47)
Back in second quantization, the spin electrons have their Fock Hamiltonian
H =0 n0 +
k nk +
{Vk ck c0 +h.c.}
(5.48)
where < n > is a parameter; thus the equations for are coupled.
No privileged spin direction exists in this problem, and one could have the
impression that n0 is equal to n0 for symmetry reasons: remarkably,
this is wrong if U is large, and the level is about half lled, because the
symmetry is spontaneously broken (see Fig. 3.3). There is a couple of ground
states, namely
n0, >> n0, ,
and another one with . The overall symmetry of the problem is
not respected by each ground states. The chemisorbed atom has a magnetic
moment. The existence of a localized spin requires strong enough U and
partial occupation; the Hartree-Fock approximation is known to overestimate
somewhat the occurrence of magnetism compared to more rened approaches.
Despite this possibility, Newns found a nonmagnetic solution for H
chemisorbed on Cu and Ni; let us consider more closely the case when
n0 = n0 = 12 n where n0 is the total population of the ad-atom.
Then, 0 does not depend on spin but according to (5.47) it depends linearly
on n0 , with 0 I for n0 0 and 0 A for n0 2. On the other
hand, the Fano formalism applied to the Fock Hamiltonian yields another
functional dependence through
n0 = 2
EF
d0 ().
(5.49)
93
n n
n n
a)
b)
Fig. 5.4. Qualitative dependence of the energy E on local spin when the virtual
level is about half lled: a) for U small compared to level width b) strong coupling
case.
(5.51)
this is conventional term meaning that we are not assuming that all quantities
be spin-independent.
4
Linear Combination of Atomic Orbitals.
94
(5.52)
k
Vk2
|Va0 |2 GS ()
k + i
(5.53)
1
, z = + i.
z 2 G00 (z)
We nd the solutions
G00 (z) =
(5.54)
z 2 4 2
.
2 2
Taking
the cut of z along the positive real axis, z = ||ei arg /2 . For
3i
4
z 2 4 2 = | 2 4 2 | exp 2 [arg( 2) + arg( + 2)] the cut is between 2 and
real axis, and corresponds to the
2 along the
band. Just
2 4 2 = i | 2 4 2 |, just below the cut, z 2 4 2 =
above
the
cut,
z
i | 2 4 2 |. The density of states is () = 1 ImG() computed on the
real axis, that is, just above the cut, and in order to have it positive, we must
choose
z z 2 4 2
.
(5.55)
G00 (z) =
2 2
Thus one obtains the semi-elliptic density of states
4 2 2
(4 2 2 ).
n() =
2 2
(5.56)
Re[G00 ] is the Hilbert transform of n() and is odd; it is linear in the band
and outside
2 4 2
, > 2.
(5.57)
Re[G00 ] =
2 2
95
( )
( C )
with C the center of the band. The two roots obtained in this way are the
bonding and anti-bonding levels of a localized surface molecule.
96
1
c |g ,
H d
(5.59)
(5.60)
2d + U
V
H(2) =
V
0
V
d
0
V
V
0
d
V
0
V
.
V
0
(5.61)
97
Fig. 5.5. The singlet density of states as obtained from Gss with 0 = 1, U = 6
and V0 = 0.1; the eigenvalues of H(2) are at = 1.023, 1.0, 0.0195, 4.004 and a
triplet at = 1. The singlet has little intensity in the high energy peak (here at
= 4); it is practically insensitive to changes in U in the low energy sector as long
as U is large. The peak at 0 = 0.0195 close to the Fermi level is the Kondo peak
due to spin ip. The lines are broadened into Lorentzians (width 0.02 |0 |).
For V = 0 the d level is singly occupied in the ground state and the
impurity has a net free spin; the deep level has a SU(2) degree of freedom.
For small V, the d level remains essentially singly occupied in the ground
state, but the transfer of spectral weight shows that it manages to interact
with the Fermi level. The only way this can occur is by an exchange of the
conduction electron with the localized one, while the impurity ips its spin.
To simplify the algebra, we can chop from the Hilbert space the uninteresting state with a doubly occupied d level, since it stands alone at high
energy, and write
d 0 V
(5.62)
H (2) = 0 d V
V V 0
on the reduced basis
{|1 , |2 , |3 } = {|d k , |d k |, |k k }.
(5.63)
2 0
d V
= cT hc = V 2 0 0 ;
we nd the transformed Hamiltonian h
0
0 d
d 2d +8V 2
(2)
. While
hence we obtain two singlets with eigenvalues =
2
98
(2)
(2)
V
+ 2 Vd , the ground state energy d 2
says that the
d
singlet is slightly lower than the triplet for small V , and a bound state
forms in which the spin is screened; the eigenvector (on the basis (5.63)
(
(2)
(+ )2 +2V 2
,
(2)
(+ )2 +2V 2
(2)
+
(2) 2
(+ ) +2V 2
,
|d k | components are the important ones. The low energy sector (singlet ground state and triplet excited level) is well represented by an eective
Hamiltonian
(5.64)
Hef f = 2JS1 S2
which commutes with the square of the total spin and describes a spin-spin
2
coupling between the two electrons. The small separation 2 FVd between
the ground state and the triplet implies that the low-energy excitation is
a spin-ip, while charge excitations are frozen. The coupling between the
conduction electron and the localized one through spin produces a singlet
(2)
level in the valence region at + .
The narrow-band model is extremely simple, but calculations on larger
systems conrm and extend the scenario. Recent exact-diagonalization calculations of Kondo clusters alloyed with mixed-valence impurities in the presence of disorder are of great interest [36] and show T=0 phase diagrams very
rich in structure.
5.3.2 Anderson Model, s-d Model and Kondo Model
Consider an Anderson model with d < 0 and U chosen such that for Vk 0
the d level is singly occupied in the ground state. At small Vk , the lowenergy subspace, say, subspace A, corresponds to singly occupied d level;
there are two high-energy subspaces, say B and B, with excitation energies
E(d1 d2 ) = d + U and E(d1 d0 ) = d . Let us use the SchrieerWol transformation ( Sect. (1.3)) with the hopping term as v; Equation
(1.86)yields the approximate renormalized interaction
(ck d )(d ck ) (d ck )(ck d )
.
(5.65)
+
Hint =
Vk Vk
d + U
d
k,k
99
and so on; if k were equal to k this would be the representation of the spin
operator in the k basis, but we are allowing o-diagonal elements. In this
way,
(kk )z = ck ck ck ck , (kk )+ = 2ck ck .
The s-d model proposed by Zener[28] for an impurity spin S in a metal reads
6
15 +
+
H = HF +
S (kk ) + S (kk ) + Sz (kk )z (5.67)
Jk,k
2
k,k
(5.68)
one can readily verify the angular momentum commutator relations in the
subspace with n n = 0, n + n = 1.
Spin-ip and the Kondo Model
Now in Equation (5.65) we consider the spin-ip (
= ) terms ignoring the
dull potential scattering terms; for instance we have a term ck d d ck =
1
+
2 (kk ) Sd which ts the s-d model. So we arrive at the Kondo model[35]
H = H0 + HK =
k ck ck +
Jk,k ( S) ck ck
(5.69)
k,k
where
Jk,k =
Vk Vk
1
1
.
+
d + U
d
(5.70)
This describes the net eect of the virtual valence uctuations; they couple
the local spin density of the conduction electrons with the impurity spin. The
d electron is reduced to its spin degree of freedom, while the charge is xed.
The k dependence is often neglected. In this case, the Kondo interaction can
be written
(5.71)
Hint = HK = J (0) (0)( S d ),
1
where (0) = N k ck is the electron eld operator at the origin. Since
(0, 0, 1) (1, i, 0)
{ }
(5.72)
(1, i, 0) (0, 0, 1)
we may also write
+
HK = J
,
.
(5.73)
100
J
( S) ck ck
N
(5.74)
k,k
with the conduction electron spin; let | denote a lled Fermi sphere,
with the energy origin such that H0 | = 0, ci and cf creation operators
for spin-up electrons near the Fermi level,
|i = ci | , |f = cf | .
One obtains the scattering amplitude Uf i = i |U |f , still an operator in
impurity spin space, by the T exp formula in the interaction picture (2.36)
Uf i = if i
i|HK ( )|f d
1
(i)2
+
d1
d2 i|HK (1 )HK (2 )|f +
(5.75)
2
The rst-order integral gives 2(if )(HK )if , that is, an energy-independent
object. The second-order contribution, introducing a complete set, reads
1
1
(2)
Uf i =
d1
d2 ei(f 1 i 2 )
(HK )f (HK )i eiE (2 1 ) ;
2
1
Uf i (i)2 2(f i )
d2 ei(E i )2 =
ei(E i )1
i(E i )
1 (HK )f (HK )i
.
i
E i
101
q i
( S)( S) =
(1 fq ) 5
q
q i
6
Sz2 + S S+ .
(5.76)
In the terms with ci in the left, the operators appear in the order cq ci cf cq ,
so q must refer to a lled state and | involve two electrons and a hole:
f |( S)cq ci | |( S)cf cq |i
E i
q
6
fq 5 2
Sz + S+ S .
q i
k,
k nk + f
Nf
nf
102
Nf
nf nf , {m, }.
Vkm c0 ck + h.c. + U
(5.78)
,
2
W
,
f
(5.80)
2 is some average of W
(). Thus, E is xed (it corresponds anyhow
where W
to a weak chemical interaction), Nf is a weak coupling limit, and we
can set up an expansion in inverse powers of Nf . We need the right combinations of conduction states that couple to the impurity, and introduce new
conduction operators with the same symmetry as the impurity m orbital6 :
1
m
=
Vmk ,k ck
(5.81)
W () k
and a combined index {m, }. Equations (5.79) yields
5
6
m , m + = , , .
Then, letting d stand for the discrete energy summation (that will become
continuous at the end)
k1 Vk1m (k2 k1 )Vmk2
m =
dm
ck1 ck2 ; (5.82)
2 (k1 )
W
m
m k1k2
now using the second of (5.79), one arrives at the useful form
H=
Nf
d
+
()c + h.c.
d W
0
nf nf + extra terms
+f c0 c0 + U
{, }
(5.83)
103
where {, } is over distinct pairs and the extra terms describe conduction
states that are not coupled to the impurity. We shall see that the ground
state can be calculated exactly for U , Nf . We start by calling
|0 the ground state of H and | the singlet state with a lled shell of k
states up to F and empty impurity states; let
0 |H|0 = E0 , |H| = E00 , E = E0 E00 .
(5.84)
d band
Fermi level
V
f
Fig. 5.6. Left: lowest state | with no occupation of the f state . Center: hopping
leads from | to a state with a single occupation of the f state. The full circle
represents a hole and the empty one a f electron: this conguration would be the
ground state for V 0 for the hole at the Fermi level. Right: further hopping may
lead to two kind of congurations: 1) a state |E, with no electrons in the f state
and an electron-hole pair in the valence band (upper panel) 2) a state |f, f , ,
with a double f hole (lower panel). For large U , however the latter is excluded, while
the coupling to |E, is negligible for large Nf . Thus, when both conditions hold,
one can solve the problem exactly.
1
|f, =
c0 | .
Nf
(5.85)
104
by the form (5.78) of the Hamiltonian, and take the scalar product. The
nonvanishing contributions are those where the creation and annihilation operators match; then, using (5.79) one performs the k summation introducing
|2 ; the m summation then introduces a Nf factor yielding
|W
Nf
1 |Vkm |2 ( k )
() Nf ;
=W
|H|f, =
Nf k
W ()
(5.87)
If we act over (5.85) with the hopping term of the hamiltonian we generate
new states:
(5.88)
H|f,
|E,
, |f, f , , ;
here,
1
|E, =
|
Nf m, E
(5.89)
has a conduction electron with energy E > EF and keeps the hole while
|f, f ,
, has two impurity electrons and two conduction holes and is ruled
out as U . Since the normalized states |E,
and |f, both bring a 1
Nf
0
here B d provisionally stands for a discrete summation over the band energies and everything is dimensionless. We write the Schrodinger equation on
a basis set
{| , |f, 1 , |f, 2 , . . .}
with eigenvectors
v = (1, a(1 ), a(2 ), . . .)
acted by the Hamiltonian matrix (recalling (5.87))
(1 ) Nf W
(2 )
Nf W
0
Nf W
(1 )
f 1
0
HM =
Nf W
(2 )
0
f 2
(5.91)
(5.92)
0
B
()a(),
dW
(5.93)
() = (E f + )a().
Nf W
Hence,
E = Nf
d
B
()|2
|W
E f +
105
(5.94)
(5.95)
(5.98)
is the inverse life time due to hopping, a the inverse root of an energy, and
so on. Typically, 0.1eV.
5.4.1 Kondo Temperature in the Spin-Fluctuation Case
E is the energy shift with respect to | , but one is more interested in the
Kondo temperature
KB TK = = f E > 0,
(5.99)
which is the correlation-induced energy gain with respect to the non-interacting
lowest eigenvalue of |f, levels. In the spin uctuation case,when nf 1, a
2 and assumes that B ; then, setting
reasonable model takes constant W
in (5.95,5.96) W = , we nd
E =
and
Nf
E f
Nf
ln |
|=
ln | |
B
nf
Nf
.
=C =
1 nf
||
(5.100)
(5.101)
= Be Nf ;
we expect the result to be roughly proportional to Nf so we introduce this
dependence, using (5.99) to eliminate E :
106
=(
(f )
( )
Nf B
Nf ln[ NB
]+ Nf
f
f
)(
)e Nf = (
)e
.
Nf
( )
f
Nf
(5.102)
B
Nf
ln(
).
Nf
(
f )
Nf
f
Nf
(5.103)
is large and negative and
|e |
Nf Nf f
)e
.
=(
(5.104)
The population of the f level can be deduced from (5.101), which says
nf
Nf
=
. Since nf is close to 1, the solution is well approximated
that 1n
f
by
1 nf = e
| |
f
f
(5.105)
i,j
0 |cm |i i|
1
|j j|cm |0 .
z E0 (N ) + H
(5.106)
One can compute the matrix elements i|z H|j without diculty and then
1
obtain i| zE0 (N
)+H |j by a matrix inversion in analogy with Andersons
procedure used above in Sect. (5.1.2). For large U and Nf one can limit
the analysis to the above approximation; the relevant states are the one-hole
states
(5.107)
|m = m | ,
such that m|m = ( ), and impurity electron -two hole states,
namely,
c
2 m2 2 1 m1 1 | .
|0m , 1 m1 1 , 2 m2 2 = 0m
Nf 1 pairs
(5.108)
107
Here, pairs sums over dierent m, channels (the normalization is to Nf 1
since the case of {m1 , 1 } = {m2 , 2 } is excluded, but 1 is negligible compared
f
(5.109)
( )
. (5.111)
0
()|2
d|W
z E Nf B zE+
f
Problems
5.1. Find by direct matrix inversion G00 () as the 00 element of ( h)1 ,
with the matrix of h from Equation (5.3). This is instructive and very easy!
5.2. Prove Equation (13.136).
5.3. Find the other elements of the resolvent matrix by the method used in
the text for G00 .
5.4. What does the Sz factor in (5.77) mean?
5.5. Prove Equation (13.136).
H =
(6.1)
110
where N is the number of electrons and A the vector potential. One also
often nds an alternative formulation, that comes directly from the relativistic
theory, namely
1
H = 2mc
c
i
N
i) =
d3 xA(xi ) j(x
Mmn am an
(6.2)
mn
where the current operator is (4.23); in the second-quantized form,
mn
runs over an arbitrary complete orthonormal set of one-body states (the set of
Hartree-Fock spin-orbitals is a convenient choice). Let |i , Ei denote the initial
state of the sample and its energy, h and k the photon and photoelectron
energies. The cross section can be worked out starting from the Fermi
golden rule
2
=
|f |H |i |2 (
h + Ei Ef )
(6.3)
h
where |f , f are the nal state of the sample and its energy. For the sake
of simplicity we shall consider the case of fast photoelectrons, when we can
neglect the post-collisional interaction between the photoelectron and the
be the N-1-electron Hamiltonian of the sample
ion left behind. We let H
f = E
f |f and write Ef = k + E
f . The
after photoionization, with H|
photoelectron state can be approximately described as a plane-wave of wave
=
(k )
|
f|am |i |2 (Ei Ef k h).
h
m
(6.4)
1
am |i .
H + i0+
(6.5)
Due to the long range of the Coulomb potential, one should use Coulomb waves,
but at high kinetic energy one has a good excuse for using plane waves instead.
2
The outgoing electron can lose energy by collisions, producing secondary electrons; this is not contained in the above simplied description; the spectrum far
from threshold when such eects are important becomes dicult to analyze.
111
112
(N O) =
|ss|, sz = 12
|ss|, sz = 1
2
(6.6)
|s|, sz = 1
(N O+ ) = |s|, |s|, sz = 0
(6.7)
|s|, sz = 1.
All these states have = |Lz | = 1, where z is the internuclear axis, but not
all are spin eigenstates. The singlet is
1
(1 ) = [|s| |s|];
2
(6.8)
|s|, sz = 1
+ |s|, sz = 0
|s|, sz = 1.
1 |s|
2
(6.9)
Let us write the total energy of the states with sz =0; since
H=
hi +
1
,
r
i<j ij
(6.10)
1
||s|
r12
(6.11)
( g ) =
| + |,
1 [| + | + | + |],
2
+
| |,
113
sz = 1
sz = 0
sz = 1
To write the nal state, let us include the unpaired 1s level (that in the initial
state was understood with the other closed-shell states). 6 states are found,
including the sz = 32 component of 4 , namely, (4 ) = | + s|. The
other components of 4 are easily obtained by the S operator and the
doublet 2 by orthogonality. In this way one explains the splitting, which is
again due to an exchange interaction.
6.1.3 Shake-up, Shake-o, Relaxation
In the nal state of core-level photoemission, a localized hole exists, producing
a eld and a polarization of the system. The polarization phenomena are
not included in independent particles theory. However, they are commonly
observed: they shift the levels towards lower binding energies. The nal state
eects show up in the spectrum also with the presence of satellite peaks.
The electron contribute to the relaxation shift, since the spectator electrons
actually are involved in some measure in the photoemission process; there
is some probability that the ion is left, in the nal state, excited. We can
adapt the independent particles theory in order to to include some correlation
eects, by using dierent orbitals for the initial and nal states. Neglecting the
energy dependence of the density of states (k ) and of the matrix elements,
the shape of the photoelectron spectrum from a deep level |c is given by
1
() = ImGc () = i|ac ( H)ac |i = i; c|( H)|i; c ,
(6.12)
where |i; c = ac |i is the N-1 electrons state that is obtained from the initial
state |i by creating the core hole but holding the orbitals frozen. We calculate
|i in the Hartree-Fock approximation for the neutral sample; under known
conditions |i; c is a single Slater determinant. Introducing the eigenstates |
and eigenvalues of H with N-1 electrons, we nd
() =
|i; c| |2 ( ).
(6.13)
However, the | eigenstates must be calculated in the presence of the corehole, that is, as determinants formed with relaxed spin-orbitals. We can obtain them too by the Hartree-Fock approximation using an excited N-1 electron conguration with the core electron missing3 . Since the overlap of determinants is the determinant of overlaps (Equation 1.5) all the many-body |
3
Unfortunately, this approximation has the shortcoming that the excited state
is not orthogonal to the ground N-1 electron state or to the lower states of the same
symmetry. This is a general drawback of the Hartree-Fock approximation.
114
and
(6.15)
320
340
360
380
Kinetic Energy (eV)
Fig. 6.1. Sketch of the satellite region of the Ne 1s photoemission spectrum (data
from Ref. ([94])). The peak in the extreme right at 383 eV kinetic energy is due
to 1s holes in the nal state; the satellites occur at lower kinetic energy because
the N e+ ion is left in an excited state. Several satellites are due to states of the
2p5 np1 2 S nal congurations, with n=3, 4, 5, . The most intense satellites are
some 20 times smaller than the main line, since Ne is a lled-shell system with low
polarizability.
115
distance. The energy of the transition between the vibrational ground states
of two electronic states of the neutral species and of the ion is called adiabatic ionization energy. The satellites are on the high binding energy side
and appear as energy losses; they correspond to vibrationally excited ions in
the nal state. The energy spacing between satellites is the vibration energy
of the ion, which diers from that of the neutral species.
When the equilibrium distances dier slightly compared to the widths of
the Gaussian wave functions, there is a high probability to end up in the
vibrational ground state of the ion, and this corresponds the most intense
peak in the photoemission spectrum. The satellites are small and fast decreasing. On the other hand, if the change in equilibrium distance is large,
the intensity of the adiabatic transition is small, the spectrum has many lines
of comparable intensities and appears almost continuous, with an intensity
maximum at the vertical ionization energy, i.e. the energy dierence between
the two potential energy surfaces at the ground state bond length. To assign
the peaks in the photoemission spectrum, an important hint comes from the
vibrational structure; this is much more intense for bonding than nonbonding
or core levels.
6.1.4 Lundqvist Model of Phonon and Plasmon Satellites
The coupling of a core electron to a vibrational mode aects the Photoemission line shape. To model the vibration, consider a single Boson mode
0
of frequency 0 =
2 with vibration coordinate x; in this way we disregard
the dierence between the initial and nal state frequencies, but such details
are readily xed if necessary. Before the photoemission event, the harmonic
potential of the vibration of the ion can be written V (x) = 12 M 02 x2 and has
a minimum for x = 0. The initial state Hamiltonian of the vibration is
H0 = 0 d d,
(6.16)
with [d, d ] = 1; the initial state |i is the vacuum, |i = |0d , and the excited
eigenstates are
1
|nd = dn |0d .
(6.17)
n!
The photo-ionization produces a sudden change in the Hamiltonian; the
potential is still harmonic, but the minimum is shifted. For a harmonic potential, the shift x x + produces a change V (x) V (x) + M 02 x
apart from a constant 2 . In second quantization, the new interaction
term proportional to x, may be written in the form
H1 = g(d + d ).
(6.18)
116
s = d , s = d ,
(6.19)
g
;
0
(6.20)
(6.21)
2
would obtain the exact result foe E. Note that the ground state |0s of H
4
is the s vacuum while |i is the d vacuum . Since
d|0s = |0s
(6.22)
(6.23)
|i|n |2 ( En ),
(6.24)
n=0
so we need the Franck-Condon factors |i|n |2 |0d |ns |2 . Let 0d |0s = C;
then, 0d |1s = 0d |s |0s = 0d |d + g0 |0s = g0 C, since the d contribution
is 0. In general,
g
1
1
0d |ns = 0d |(s )n |0s = ( )n C.
n!
n! 0
4
(6.25)
117
C = e 2 a , a =
Hence,
g2
E
=
.
2
0
0
an
( E n0 ).
n!
n=0
L() = ea
(6.26)
(6.27)
The probability of n excited bosons Pn = an! ea follows the Poisson statistics, reecting the statistical independence of the non-interacting bosons, and
the average number of excited bosons is
n = a = 2 .
(6.28)
dL() = ea
an
(n0 + E) .
n!
n=0
an
an1
n0 =
a0 = Eea .
n!
(n
1)!
n=1
n=1
The center-of-mass of the line remains where it would be for g=0; the threshold is at lower binding energy, at a0 below the center-of-mass. Fourier transforming L() one nds the correlation function
dL()eit = i|eiHt |i
2L(t) =
= ea
Thus,
an ia0 tin0 t
e
.
n!
n=0
(6.29)
i|eiHt |i = eC(t) ,
(6.30)
(6.31)
where
Note the characteristic exponential at the exponent. For strongly coupled,
slow modes, the a 1 case is relevant. Formally, we let 0 0 with g 2 = a02
nite. Then, eiHt e
g2 t2
2
2
1
L() = e 2g2 .
g 2
(6.32)
118
The relaxation energy diverges in the Gaussian limit, which is a serious overestimate; however the Gaussian line shape is often a good approximation
for phonon broadened core levels in solids and the width provides a sensible
measure of the electron-phonon interaction, although the Gaussian should be
convolved with a Lorentzian (producing a so called Voigt prole) to account
for the core hole lifetime. The result
lim ea
2
an
1
( + a0 n0 ) = e 2g2
n!
g 2
n=0
(6.33)
(6.34)
cm cm + Hr
(6.35)
m,
comprises a spin-up deep hole term for the primary hole, cm creates a
valence hole with magnetic quantum number m; Hr is the hole-hole repulsion Hamiltonian with screened direct and exchange matrix elements
U (m1 , m2 , m3 , m4 ). The crystal Hamiltonian HS and the atom-crystal hopping term Hint are one-body contributions that can be chosen according to
any model deemed appropriate. Finally,
Hf =
k ck ck
119
(6.36)
k,
describes the free particle continuum above the vacuum level; in the ground
1
|m (1)m (2)
r12
(6.37)
(6.38)
with two holes localized on the valence orbitals of the same atom that was
ionized initially. At this point Hf may be forgotten.
By the Fermi golden rule one nds[76] that the spectrum is proportional
to the free-electron density of states and to S(Ei Ek , k), given by
Amn (k)Apq (k)Dmnpq ()
(6.39)
S(, k) =
mnpq
where the local density of states (LDOS) for the two nal-state valence holes
is
h Hr H1 )|mn ;
(6.40)
Dmnpq () = pq|(
here, H1 is the sum of all the one-body terms of H Hf . The dependence
on the angles of k can be taken care of by multiple scattering techniques, in
the spirit of Ref. [86], but here we concentrate on the dependence.
The fact that the local density of states appears, rather than the band
density of states as in the Lander theory, is one of the main points of Ref.[76]
and qualies Auger spectroscopy as a local probe of valence states. There
are two trivial special cases. In the atomic limit Hint = 0, and HS = 0;
then the Hamiltonian is diagonal in the L-S or |LSM MS representation;
the spectrum consists of unbroadened multiplet terms. The other simple
limiting case is the non-interacting (Hr 0) case; density of states ma(0)
trix Dmnpq () is readily worked out in terms of the one-body local denh H)|n = nm () using their transforms;
sity of states mn () = m|(
here we dene the correlation functions Dmnpq (t) = pq|eiHt |mn and
mn (t) = m|eiHt |n , both with diagonal elements equal to 1 for t=0.
Indeed,
(0)
(6.41)
120
1
|00 = i[I (0) () iD(0) ()],
H0 + i
with
I (0) () =
d
D(0) ( )
.
(6.42)
(6.43)
One obtains
() =
(0) ()
,
1 + iU (0) ()
(6.44)
iii)
(x + a)(c x) = (a + c) [(x + a) (x c)]
D() =
where iii) comes from ii) and (x) = 1 (x). One nds the triangularshaped result
D0 () = (2 ||)(
||
1
2 ).
2 4
121
(6.48)
0
As an alternative route to (6.48) we can use (t) = sin(t)
t , D (t) =
2
sin(t
, and as this is symmetric we can compute the real Fourier transform
t
by calculating the cosine transform:
0
2D () =
dt cos(t)D0 (t).
(6.49)
I () =
dyD0 (y)
y
1
+ 2
2 42
=
log |
| + 2 log |
|.
2
2
4
2
(6.50)
i(0) (z) =
z + 2
z
1
z 2 42
log(
)+ 2 log(
) = I 0 ()iD0 (). (6.51)
2
z 2
4
z2
122
D()
-2
D()
-2
Fig. 6.2. For the rectangular band model, the two-hole line shape is triangular
for U = 0 (left panel,dotted), and is shown here broadened with = 0.01. With
increasing U the maximum becomes sharper and shifts to higher , as shown for
U = 1.1 (left panel,solid). At U = 1.8 a well developed split-o state develops
(right panel, dotted). The solid line in the right panel shows the result for U = 3
when most of the intensity is in the two-hole resonance, but a residual band-like
continuum can still be seen.
123
ionization occurs in the Ti 2p level, but the hole has little chance to Auger
decay via intra-atomic transitions since the Ti ion (nominally, T i4+ ) has
no valence electrons left. So, the electrons must come from the surrounding
closed-shell O ions. One can consider electron transfer to the empty Ti
valence shell and/or direct inter-atomic decay. When two electrons come from
dierent O ions no desorption occurs; indeed, desorption takes several phonon
periods (of the order of 1013 s) and bond healing by hole delocalization
should be much faster. However, there is some chance of two holes in the
same O. In this case, a two-hole resonance forms and lives long enough to
allow the O+ ion to escape. The generalization of the theory to partially lled
bands is discussed in Section 12.3.
124
Fig. 6.3. Sketch of the Au N6,7 O4,5 O4,5 Auger spectrum (Ref. ([153])) compared
with theoretical proles (see text). The solid line is a drawing of the experimental
prole, on a binding energy scale, where 0 corresponds to both holes from the Fermi
level; the theory used the Slater integral F 0 , or equivalently one of the U values,
as an adjustable parameter. The dotted line is the best shift to the theoretical
line shape obtained with U (1 G4 ) = 3.4eV, but appears to be shifted compared to
experiment. By increasing U one can adjust the line position, but the line shape
agreement is lost (dashed line). To x this problem, one must account for the o-site
interactions (see below).
G(R1 , R2 , R3 , R4 , z) = R1 R2 |
1
|R3 R4
zH
(6.52)
R,
Here, belongs to the lattice of relative motion, which we call the lattice,
isomorphous to . For V = 0, the problem is trivial because H = H + H ;
then the time evolution operator factors and the non-interacting Greens
function that we denote by a small g is given by
g(R1 , R2 , R3 , R4 , z) = g (1) (R1 , R3 , z) g (1) (R2 , R4 , z)
(6.54)
(1)
125
|; Q =
|Rcm , Rcm + =
e
|R, R +
2
2
N
N R
Rcm
(6.56)
where the integration variable was shifted using (6.55). Moreover,
1 iQ[R+ 12 ]
|R, R + =
e
|; Q .
N Q
(6.57)
One can immediately check that this {|; Q } set is a correctly normalized
basis for two particles.
Fourier Transforming the Relative Motion
The trick is
eiQR
eiQR
|R, R + =
(, )|R, R +
N
N
R
R
where we can use (, ) =
1
N
1
N
but
R,
3
2
3
2
eiq
eiq
(6.58)
eiq( )
.
N
eiQR+iq |R, R + =
R,
ei[Qq]R+iq[R+ ] |R, R + ,
(6.59)
R,
R,R+
eiQR
1 iq
|R, R + =
e
|Q q, q ,
N
N q
R
(6.60)
where
|Q q, q =
1
N
(6.61)
R,R+
exp( 2i Q ) iq
e
|Q q, q .
N
q
(6.62)
126
Projected Hamiltonian
In the |, Q basis,
H=
H Q , H Q = H0Q + H1Q ,
(6.63)
(6.64)
while
H0Q =
,
Q
|, Q W
Q, |;
(6.65)
Q
we determine W
using (6.62): one nds
Q
2 Q( )
W,
= e
i
1 iq( )
e
[(Q q) + (q)].
N q
(6.66)
1 iQ[R+ 12 ( )] Q
1
G (z),
|R, R + =
e
zH
N
(6.67)
1
|Q, .
z HQ
(6.68)
Q
2 Q[ ]
g,
(z) = e
i
q
eiq[ ]
.
z (Q q) (q)
(6.69)
(6.70)
1
1
1
1
1
1
=
+
V
+
V
z HQ
z H Q z H0Q
z H0Q
z H0Q
z H0Q
1
1
1
1
=
V +V
+
V
.
(6.71)
Q
Q
Q
zH
z H0
z H0
z H0Q
127
Expanding the inverse operator inside the curly brackets and summing the
geometric series, we obtain
1
1
1
1
1
=
+
V
.
1
Q
Q
z HQ
1
V
z H0
z H0
z H0Q
zH Q
(6.72)
We take matrix elements of (6.72) on the |; Q basis and we nd, adopting matrix notation on the lattice, with V diagonal,
5
61
V gQ.
GQ = g Q + g Q 1 V g Q
(6.73)
This is the solution. The size of the matrix that must be inverted depends
on the range of the potential. In the zero-range case of the on-site potential
(Hubbard Model with local interaction U ) we set = , and write
g Q (z) =
q
1
.
z (Q q) (q)
(6.74)
Then, Equation (6.73) is scalar and the solution can be simplied to read:
GQ =
gQ
.
1 U gQ
(6.75)
(6.76)
with a small positive constant acting on an interacting system with Hamiltonian H; here f is a parameter and V a time-independent operator. We
want the wave function (t) that reduces to the ground state eiEt g for
r . In the interaction picture,
128
(6.77)
with the starting condition I (t) g in the remote past. In rst approximation, we write (setting h = 1)
i
(6.78)
1
V g .
zH
(6.79)
In the Schr
odinger picture,
(t) = eiEt g + f eizt
1
V g .
zH
(6.80)
Photoemission
e
Now let V = mc
i A(xi ) pi (summed over electrons) the operator which
produces photoemission. In second quantization,V = p ,d (p , c)cp cd where
deep a electron is annihilated and a photoelectron created, and (p , c) are the
matrix
elements. We assume that H includes the photoelectron kinetic energy
Tpe = p p np but no post-collisional interaction with the photoelectrons.
The photoelectron current is
Jp =
d
(t)|np | (t) .
dt
(6.81)
in the above approximation, the current is a quadratic response in the perturbation V (which is logical, since it is a d.c. response to an a.c. perturbation.)
Since np g = 0,
np = f 2 e2t g |V
1
1
np
V |g .
H zH
(6.82)
1
1
np
V |g .
z H z H
(6.83)
Hence,
Jp = 2f 2 e2t g |V
Now since [np , H] = 0, we develop:
Jp = 2f 2 e2t
dd
letting 0,
1
1
cd |g
H zH
(6.84)
Jp = 2f 2
dd
cd |g
|z p H|2
129
(6.85)
where the photoelectron energy p has appeared (up to now, Tpe was part
of H, but now we use H for the system Hamiltonian without Tpe .) This is
actually
Jp = 2f 2
(6.86)
dd
so the result depends on the one-hole density of states matrix dd () of the
system. This formalism generalizes the elementary theory of photoemission,
but oers the possibility of a clear enhancement of the understanding of the
Auger eect.
6.4.1 One-Step Theory of Auger Spectra
Let k label the Auger electron; for the current Jk =
like above
Jk = 2f 2 e2t g |V
d
dt (t)|nk | (t)
we nd
1
1
nk
V |g , (6.87)
(H + W + T ) z (H + W + T )
where now, with a slight change in notation, we made explicit the operator
(6.88)
W =
Whh dk cd ck ch ch + h.c.
producing the Auger transitions; ck and cd create the Auger electron and the
deep electron, respectively, and the annihilation operators create the pair of
nal-state holes; also, T = Tpe + TA the kinetic energies of photoelectrons
and Auger electrons. Using
1
1
1
1
=
+
W
zH W T
zH T
zH T zH W T
and
1
1
1
1
=
+
W
zH W T
zH T
zH W T zH T
1
V |g = 0 one is left with
since nk zHT
1
1
W
nk
H W T z H T
1
1
W
V |g
zH T zH W T
Jk = 2f 2 e2t g |V
(6.89)
130
1
1
nk
= nk (E + H T ).
H T zH T
So, we reach the general expression of the one-step description of the Auger
eect by Gunnarsson and Sch
onhammer [22]:
1
W nk
z H W T
1
V |g .
(E + H T )W
zH W T
Jk = 2f 2 g |V
(6.90)
The most important terms are diagonal in the deep hole; writing V explicitly
and replacing T by TA + p , where p is the photoelectron energy,
Jk = 2f 2
p,d
(E + H T )W
1
W nk
z H W p T A
1
ad |g .
z H W p T A
(6.91)
Problems
Part II
G is closed, i.e. a G, b G = ab G.
The product is associative : a(bc) = (ab)c.
an identity e G such that ea = ae = a, a G.
a G, a1 that is every element has an inverse, such that a1 a =
aa1 = e.
134
ab
x
ab
x
=
and
=
, one nds the condition
y
cd
y
cd
ab
Det
= 1, so one is left with SL(2). Usually one proceeds the other
cd
way: one knows a Group H, discovers new operations and so builds G.
Let H be a subgroup of G; for A G consider the set
H(a)
= {aha1 , h H}.
Since ah1 a1 ah2 a1 = ah1 h2 a1 , this is a subgroup of G, the conjugate
subgroup with respect to a. Occasionally, it may coincide with H itself (as
a set, not element by element), in which case we write aH = Ha. If a
G, aH = Ha, then H is called invariant subgroup of G, or normal divisor of
G.
L
h
135
(7.1)
= |R | = |R1 | = ei | ;
R
(7.4)
(7.5)
| D (S).
(7.6)
136
| D (S) =
| D (R)D (S) =
| D (RS)
(7.7)
with
D (RS) =
D (R)D (S).
(7.8)
thus, the matrices are multiplied like the operators and in the same order,
providing a representation of the Group.
In simple cases, all the symmetry operators commute; then we can diagonalize them simultaneously with the Hamiltonian matrix |H| . We can
diagonalize R, then with the new basis diagonalize S, and continue until we
have a new basis labeled with all the symmetry-related quantum numbers.
This breaks up the Hilbert space into several orthogonal subspaces. Then,
we can diagonalize H separately in each subspace. In such cases,we get the
maximum simplication from symmetry without Group Theory. The names
of Groups are just nicknames for the symmetry of the problem.
Blochs Theorem
In band structure calculations one solves
2
p
+ V (x) (x) = (x)
2m
(7.9)
ipti
(7.10)
(7.11)
with
uk (x) = uk (x + ti )
(7.12)
137
(p + k)2
+ V (x)]uk (x) = k uk (x)
2m
(7.13)
Water Molecule
Suppose want to calculate the molecular orbitals of H2 O in a LCAO model
neglecting overlap. Putting the molecule on the xz plane, with z as the molecular axis, we see that it belongs to the C2v point Group, with the operations
C2 : (x, y, z) (x, y, z); (xz) : (x, y, z) (x, y, z); (yz) : (x, y, z)
(x, y, z) We can form the Multiplication Table:
E
C2 (xz) (yz)
C2
E (yz) (xz)
(xz) (yz) E
C2
(yz) (xz) C2
E
the Group is Abelian and the square of each operation is the identity E.
Therefore the eigenstates of H will be even or odd under any of the operations.
2 | ,
From an arbitrary basis {| } we can generate 2 bases |C2 = 1C
2
then 4 bases |C2 (xz) =
1(xz) 1C2
| ,
2
2
| .
2
2
2
I
1
1
1
1
C2 xz
1 1
1 1
1 1
1 1
yz
1
1
1
1
g=4
z
xy, Rz
x, Ry
y, Rx
Such are the possible symmetry types of the solutions,that have conventional names shown in the rst column.
The C3v Group (NH3 )
The C3v (or 3m) Group is the symmetry Group of an equilateral triangle (or
the Group S(3) of the permutations of 3 objects). Let the vertices be labeled
(a,b,c). The operations are a C3 rotation, its square (or inverse, which is the
same) and three vertical planes a , b , c through the center and the vertices.
The sense of rotation is arbitrary and we may choose C3 as the operation
(a, b, c) (c, a, b); clearly a (a, b, c) (a, c, b). Next, we need a convention
138
about the multiplication of two operation. a C3 (a, b, c) = a (c, a, b) The commonly adopted one states that the result is (c, b, a), that is, a keeps the rst
entry xed, although the rst operation sent a elsewhere. In other terms, it is
the position of symmetry elements that matters, while the label can change.
In this way, one notices that a C3 = b , but C3 a = c , thus the Group
is not Abelian. All the information about the abstract Group is in the multiplication table.
E C3 C32 a b c
C3 C32 E c a b
C32 E C3 b c a
(7.14)
a b c E C3 C32
2
b c a C3 E C3
c a b C3 C32 E
The rearrangement theorem holds:
Theorem 2. Each line and each column contain all R G.
This follows from the denition of Group. In any line or column there are g
elements, and all are distinct since for instance C3 R = C3 S R = S. Every
operation does a permutation of the vertices. In C3v the converse is also true,
so C3v is isomorphous to S(3). Already in C4v the 8 operations are fewer
than the 24 permutations of 4 objects. The Cayley theorem states that
Theorem 3. Any group of order NG is isomorphous to a subgroup of S(NG ).
Z3 = {E, C3 , C32 } C3v is an invariant Abelian subgroup. A Group G is
called simple if it does not have invariant subgroups. So, Z3 is simple, but
C3v is not. A Group G is called semi-simple if it does not have abelian
invariant subgroups. So, Z3 is semi-simple, but C3v is not.
It is interesting to nd representations of Z3 using the multiplication table
above. Besides the trivial one, with all the operators represented by 1 (the
2i
A1 representation below in Table 7.1 ) one can represent C3 by = e 3
and C32 by the third cubic root of 1, namely ; in a third representation,
C3 , C32 . Thus, there are two complex conjugate representations.
The p orbitals of an atom centered at the origin behave dierently under Z3 : the z orbital, that we assume parallel to the rotation axis, is not
transformed and belongs to A1 . The combination x + iy is multiplied by
and x iy is multiplied by ; thus the (x,y) pair is a basis for the conjugate
representations. This is reported in the last column of Table 7.1 )
The Quotient Group
Let H be a subgroup of order NH of G and g G. the sets
Hg = {hg, h H}, gH = {gh, h H}
are respectively right and left cosets. Both have exactly NH elements; for
instance, h1 g = h2 g multiplied on the right by g 1 becomes h1 = h2 .
=e
z
139
2i
3
(x, y)
For example, in C3v , using Z3 and a one forms the right coset
{E, C3 , C32 }a = {a , b , c }.
Using b or c one nds the same coset, while a rotation gives Z3 . Actually,
we have split C3v in two subsets of 3 elements each as follows:
C3v = Z3 + Z3 a .
(7.15)
140
nC
T,
(7.18)
T C
that commutes
all G, that is X G, C = X 1 C X. Indeed,
nwith
C
1
1
X C X =
T X has nC terms conjugated with T and such
T C X
terms are all dierent because X 1 T X = X 1 T X T = T . So, we have a
quantum number for each class.
For C3v , the Dirac characters are
E = E, R = C3 + C32 , = a + b + c .
These are simultaneously diagonal and their eigenvalues E = 1, R , (except the rst, which is trivial) are useful wavefunction labels. They are not
independent, but occur in combinations. Using the multiplication table,
2
this term is rather strange, since an element can be conjugated to any number
of elements.
(7.19)
hence
141
= 2 + R R =
2
1
2 = 3 + 3R 2 = 3 + 3R .
(7.20)
(7.21)
3
-3
0
(7.22)
(7.23)
Fig. 7.1. The chosen geometry, with the 1 reection that changes x into -x.
orbitals or any pair of functions transforming like (x,y) will do the same):
10
c s
1 0
D(E) =
; D(C3 ) =
; D(a ) =
(7.24)
01
s c
0 1
1
2
3
with c = cos( 2
3 ) = 2 , s = sin( 3 ) = 2 . These are the generators, i.e., the
others can be obtained by multiplication:
c s
c s
; D(b ) = D(a )D(C3 ) =
;
D(C32 ) = D(C3 )2 =
s c
s c
c s
D(c ) = D(a )D(C32 ) =
.
(7.25)
s c
142
1 i
2
2
i
1
2
2
2i/3
3 ) = U D(C3 )U = diag(e
is unitary (U U =unit matrix), D(C
, which
2i/3
, e
)
0
1
2
143
is the same for all D matrices in the given class C, since the Tr operation
is invariant under unitary transformations like conjugation. (C) is called
character, and is related to Diracs character. The matrices of the C are
constants3 , hence, recalling the denition of C
T rC = mC = nC (C).
Thus,
(C) =
m
C .
nC
(7.28)
(7.29)
The characters (C) are very useful, as we shall see, and are tabulated for
the point Groups (See Appendix II; note that these tables are square (the
number of irreps is equal to the number of classes).). In particular T rE = m
can be read o the Tables.
Actually, the tables were computed (and can be found for new Groups
when necessary) from the multiplication Table; one rst deduces the classes
and c as in the above example. One does not know m a priori but shall
derive below
NG
(7.30)
m2 = 2 .
C
C nC
With this result, the reader is already able to calculate the character tables.
7.2.3 Schurs lemma
We have shown above that the matrices of Dirac characters and of H within
an irrep are multiples of the identity matrix. Due to the importance of this
crucial point Ill present a second proof, the traditional one based on Shurs
lemma. The two arguments enlighten dierent facets of the problem; it will
be apparent below that this is just a matter of algebra.
Lemma 1. Any m m matrix M commuting with all the representative matrices D(Ri ), 1 = 1, NG ) of an m-dimensional Irrep must be proportional
to the Identity matrix Imm .
3
144
We rst prove the lemma for a 2 2 diagonal matrix, then generalize to any
matrix of any size.
Proof for a diagonal matrix
a11 a12
a21 a22
0
a12 (d2 d1 )
00
=
.
[A, ] =
0
00
a21 (d2 d1 )
, and
Now, if d2 were
= d1 , this would require a diagonal A. Since one cannot
diagonalize simultaneously all the D(Ri ), 1 = 1, NG ), we conclude that
d2 = d1 .
In the 33 case, we let M = = diag(d1 , d2 , d3 ), and the same reasoning
leads to
0
a12 (d2 d1 ) a13 (d3 d1 )
0
a23 (d3 d2 ) .
[A, ] = a21 (d1 d2 )
a31 (d1 d3 ) a32 (d2 d3 )
0
If all the di are dierent, this vanishes only for diagonal A. If d1 is dierent
from the other diagonal elements, the a11 element must be one block, separated from the rest, which cannot be true for all the representative matrices
in an irrep. This is formally generalized formally to matrices of any size with
the conclusion that = Imm for some .
Proof for a Hermitean M
Assuming M hermitean we know that by a unitary transformation U it can
be diagonalized U M U = = Imm and actually this implies that
M = Imm in any basis.
Proof for any M
We can do without the assumption that M is Hermitean. Indeed, H1 =
M + M and H2 = i(M M ) are Hermitean at any rate. Now, we
show that M also commutes, for, by taking the Hermitean conjugate of
D(Ri )M = M D(Ri ), one nds M D(Ri ) = D (Ri )M dag . Now we multiply on the left and on the right by D(Ri ) and since it is unitary we nd
D(Ri )M = M D(Ri ). Then H1 and H2 also commute, and are constants.
Therefore, M = 12 (H1 iH2 ) = Imm ,
q.e.d.
7.2.4 Continuous Groups
Many Groups of fundamental importance in Physics are Lie Groups. They
are continuous (elements can be labeled by parameters) and continuously
145
vector directed along the axis and with length equal to the angle of (say,
counterclockwise) rotation ; this corresponds to a sphere of radius where,
however, each point of the surface is equivalent to the opposite one. All the
rotations with the same || belong to the same class. the angular momentum
J
J
|Jm Dm
m ().
(7.31)
m =J
146
d is higher than m. The most popular case occurs in the Schrodinger theory
of the H atom, when energy depends only on the principal quantum number
n and orbitals have a n2 fold degeneracy, while the spherical symmetry only
allows the energy to be independent on magnetic quantum number. In such
cases, one cannot produce the 2p orbitals by rotating the 2s one. In such
cases, one speaks of accidental degeneracy.
It is possible that by accident two states unrelated by symmetry come so
close in energy to appear degenerate in low resolution experiment; however
a mathematically exact degeneracy with no symmetry reason is miraculous.
Simply, one was unaware of using a Subgroup of the actual Group, because
some symmetry had still to be discovered, as in the following examples.
Hydrogen Atom
The n2 fold degeneracy of the non-relativistic H atom Hamiltonian
H=
k
p2
2m r
(7.32)
1
r
potential.
The fact that the force F = k rr3 is central explains the conservation of
d 1
dt r
r
p L
k .
R=
m
r
r2
d
v
r (
r
v)
r
=
.
dt r
r3
(7.33)
(7.34)
L
The x component of the numerator is (x2 + y 2 )vx x(xvx + yvy ) = m
y.
Developing in this way, one nds
L
d
r
=
(y, x, 0).
dt r
mr3
(7.35)
p L
L
d
kL
= (Fy , Fx , 0) =
(y, x, 0).
dt m
m
mr3
(7.36)
147
r R
r . Since it is conserved, R is pinned at the aphelion-perihelion
4
direction and this is why the orbits are closed5 .
Quantum Runge-Lenz Vector
The quantum Runge-Lenz vector
p L L
p
r
R =
k .
2m
r
(7.37)
commutes with H; the calculation can be carried out like above, with
replaced by hi [H, A] . Hence,
dA
dt
HLM = ELM H R LM = E R LM ;
but one can check that R does not commute with L2 , so R LM does not
belong to L and the energy must be L independent. More details and the relation to the O(4) Group worked out by Pauli can be found on Schis book[25].
In the relativistic case, the Dirac-Coulomb Hamiltonian HDC commutes with
the Biedernharn-Johnson-Lippman (BJL) pseudoscalar operator
i
Ze2
r
2
(7.38)
K
(H
mc
)
5
DC
2
mc
c
r
where K = L + h
, =
, = 4 , 5 = 1 2 3 4 . This is
the reason why the levels do not depend on the sign of K and for instance
the pairs (2s1/2 , 2p1/2 ), (3s1/2 , 3p1/2 ), (3p3/2 , 2d3/2 ) are degenerate in Diracs
theory.
B=
0110
1 0 0 1
h=
1 0 0 1
0110
(7.39)
148
-2 2 = ( 12 , 21 , 12 , 12 )
0 0,1 = ( 12 , 0, 0, 12 )
0 0,2 = (0, 12 , 12 , 0)
2
2 = ( 12 , 12 , 12 , 12 )
and the degeneracy is expected because the irrep E of C4v has to be represented. The eigenvalues come in pairs (||) since this graph is bipartite.
a)
3
2
b)
Fig. 7.2. llustrative example of dynamical symmetry. a) square tight-binding cluster b) deformed cluster with heavy lines denote doubled matrix elements. The symmetry Group of b) is isomorphic to the Group of the square.
0120
= 1 0 0 1.
h
(7.40)
2 0 0 2
0120
The deformed cluster has much less geometrical symmetry but as much dynamical symmetry as before. Was the degeneracy removed? No. The rst
and last lines are equal and the other two are proportional, thus two zero
has the following spectrum:
eigenvalues are still there. Indeed, h
1
1
10 ( 2 , 10 , 25 , 12 )
0
( 12 , 0, 0, 12 )
0
(0, 25 , 15 , 0)
10 ( 12 , 110 , 25 , 12 )
There is accidental degeneracy. One symmetry element is the reection
0
0
=
0
1
149
001
1 0 0
0 1 0
000
0010
0 1 3 0
1 0 0 0
3 0 0 1
S=
(7.41)
0 0 0 1 S = 10 1 0 0 3
0100
0 3 1 0
Note that S 4 = 1 and S does not produce a permutation of sites but
is a genuine generalization. Moreover S mixes 0,1 and 0,2 , does not commute with and explains the degeneracy. The deformed problem has a lower
geometrical symmetry, but actually is still C4v symmetric because of a hidden dynamical symmetry. More dynamical symmetries will be discussed in
Sect.17.0.6.
(i)
RG
(j)
D (R) D (R) =
NG
mi ij .
(7.42)
The GOT is the central result of Group representation theory, and I prepare
the proof by a few remarks.
Remark 1: I am going to present the proof for discrete Groups, however
the GOT extends to continuous ones. The following orthogonality holds for
the Wigner matrices[23]
8 2
J
a,a b,b J,J
()DaJ b () =
(7.43)
dDab
2J + 1
2
2
where , , are Euler angles and d = 0 d 0 sin d 0 d.
150
E
10
01
+
C3
1
3
2
2
3
1
2
2
,+
C32
12
23
3
2
12
c
,
a + b , +
,
1
3
1
3
1 0
2
2
2
2
3
3
0 1
1
1
2
(7.44)
Note that C32 is the inverse of C3 and its matrix is D(C3 )T . More generally,
the operators are unitary and
(i)
(i)
D
(R) = D
(R1 ).
(7.45)
(7.46)
For i
= j, i and j are orthogonal because they belong to dierent eigenvalues of at least some c . The invariant operator O commutes with the
Dirac characters, so O| j has the same characters as | j . For i = j the
matrix {O } represents an operator that commutes with everything, so it
must commute with all the D matrices; so Schurs lemma applies.
For example, a spherically symmetric potential has vanishing matrix elements between states of dierent L and within L it has vanishing matrix
elements between states of dierent ML ; the diagonal matrix elements are
independent of ML and depend on L.
Remark 4:
Theorem 7. Consider an arbitrary operator .
O=
R1 R
(7.47)
RG
is an invariant operator.
This follows
immediately from
the rearrangement theorem: T G, OT =
T T 1 RG R1 RT = T RG (T R)1 (T R) = T O.
Remark 5:
151
(example of GOT)
Let us form 6-component lists with the elements of the D matrices; if we
treat them like vectors and compute the norms we nd 6 for the irreps with
m=1 and 3 for those with m=2.
element
11 of irrep
12 of irrep
21 of irrep
22 of irrep
irrep A1
irrep A2
list
1 1
E (1,
21 ,
12 , 1,
, )
2 2
3
3
3
3
E (0,
,
,
0,
,
)
2 2
2
2
3
3
3
3
E (0, 2 , 2 , 0, 2 , 2 )
E (1, 21 , 12 , 1, 21 , 21 )
(1, 1, 1, 1, 1, 1)
(1, 1, 1, 1, 1, 1)
of squares
3
3
3
3
6
6
(7.48)
RG
(j)
|j |D
(R),
i |R =
(i)
i |D
(R) ,
we nd
(i) ||(j) =
(j)
(i)
i ||j D
(R)D
(R) = O(i)ij .
(7.49)
RG
To prove the GOT we must eliminate i ||j . Now we take advantage
of the total freedom to choose as we like. If we assume that for some pair
of components, and , say ,
= |i j |,
(7.50)
(j)
(i)
D (R)D
(R) = O (i)ij .
(7.51)
152
(7.53)
RG
(7.54)
RG
(i)
(j)
D
(R) D
(R) =
RG
NG
ij .
mi
(7.57)
In particular,
(7.58)
|(i) (R)|2 = NG ,
153
(7.59)
RG
which gives a method to verify that a given representation is indeed irreducible; if we mix two irreps i and j with coecients
ni and nj , then
5
6
(R) = ni (i) + nj (j) and the r.h.s. becomes NG n2i + n2j + 2ni nj ij .
Rewriting the LOT in terms of classes,
nC |(i) (C)|2 = NG ,
(7.60)
C
mi C
nC
we obtain
7
8
8 NG
mi = 9 2 .
C
(7.61)
C
nC
A reducible representation can be reduced to block form and the block of the
i-th irrep appears ni times along the diagonal; on any basis, the trace is
nj (j) (R).
(R) =
j
which implies
1 (i)
(R) (R).
NG
(7.62)
1
nC (i) (C) (C).
NG
(7.63)
ni =
Also,
ni =
(i)
(j)
The LOT
C (C) (C) = NG ij , so the vectors with compo becomes:
nents NG (i) (C) are also orthogonal. They have component for each class
and there is one of them for each irrep. Thus,the number of irreps does not
exceed the number of classes. Actually one can prove that they are equal, that
is, the character tables must be square [156]. This is related to the second
character orthogonality theorem
i
(i) (C)(i) (C ) =
NG
nC CC ,
(7.64)
154
that is,
(j)
(i)
D
(R) D (R) =
RG
i
D
(R) Rj =
mi (i)
D (R) Rj = ij j .
NG
NG
ij
mi
(7.65)
Hence
i
=
P
mi (i)
D (R) R
NG
(7.66)
i
is a generalized projection operator6such that P
= |i i |. In other terms,
i
P
j = ij i .
(7.67)
i
The diagonal P
operators are all what we want, and require only the
diagonal elements of the D matrices. Let = j c(j, )j belong to a
reducible representation:
i
= c(i, )i
(7.68)
P
is projected on i , unless c(i, ) = 0 (if the choice was unlucky never mind,
we shall succeed with another of the reducible set). To transform the {}
set to the new {j } basis we just need to know how every R G operates on
(i)
(i)
the old set and the diagonal elements of the D (R) matrices. The D (R)
matrices are properties of the abstract Group,and we know them in principle:
they are mi mi matrices and having xed which operators R we can and wish
to have diagonal, we have enough relations from the multiplication table to
build them; in the most common problems this is just a matter of geometry to
determine how the components of a point or other simple functions transform.
Moreover, the matrices are not really needed to project to the new basis.
Indeed,
i
P
=
c(i, )i .
(7.69)
In the combination on the r.h.s. all functions belong to the irrep i, so the
l.h.s. is the projection operator,
6
i
i
The name is deserved since the diagonal terms are such that (P
)2 = P
.
P (i) =
i
P
=
i (R) R.
155
(7.70)
The characters are enough to build P (i) and to obtain P (i) which belongs to
irrep i. Repeating this with enough , one builds a basis for the irrep. These
bases are typically small sets, where one can easily orthonormalize; usually
one readily performs the unitary transformations needed to have as many
diagonal operators as possible, in order to increase the number of quantum
numbers.
m2i = NG .
(7.71)
001000
1 0 0 0 0 0
0 1 0 0 0 0
,
D(C3 ) =
0 0 0 0 1 0
0 0 0 0 0 1
000100
156
while D(E) = Diag(1, 1, 1, 1, 1, 1). This regular representation exists for any
G; (E) = NG while for all the other classes = 0 since no other element
can have 1 on the diagonal (RX =
X R = E ) . Irrep i is contained in
the regular representation ni = N1G R i (R) (R) = N1G i (E) (E) = mi
times. Reducing the representation, we shall have on the diagonal mi times
the mi mi block, and the length of the diagonal is NG . This proves the
Burnside theorem.
Problems
7.1. Prove Equation 13.136.
7.2. Build the character table of C4v .
7.3. Build the projection operator for the irrep E, componet y, of C4v .
158
2 1 1
P (E) =
(E) (R)R = 1 2 1 .
(8.1)
R
1 1 2
The that one obtains depends on the s orbital; if we start from s1 , which
2 s3 which is
is v (1) invariant, we get (upon normalization) 2 = 2s1 s
6
also v (1) invariant. This is the x component, which is invariant under the
exchange of 2 and 3. The orthogonal function is odd under such an exchange
3
. One can get this result by the projection operator. Starting
and is 3 = s2s
2
1 s3 which is not orthogonal to 2 but has
from s2 , one obtains 3 = 2s2 s
6
the orthogonal component 3 . Alternatively, one can obtain those results by
projecting directly on the x and y components of the irrep E, by
(E)
Pxx
=
2 (E)
Dxx (R) R.
6
(8.2)
2
2
2
and 3 = 1 C3 = 23 12
. Thus, one
as 2 = 1 C3 =
3
1
2
2
2
2
obtains
2 1 1
1
(E)
Pxx
= 1 12 12 .
(8.3)
3
1 12 12
In a similar way,
0 0 0
= 0 12 21 .
0 12 21
(E)
Pyy
(8.4)
Thus, we have determined the symmetry orbitals; 2s, 2pz , 1 belong to A1 and
the pairs (2px , 2py ), (2 , 3 ) belong to E. The 7 7 determinant is broken
into a 3 3 determinant for the A1 basis and a pair of 2 2 determinants
for the x, y components of E; moreover, by the Schur lemma, the latter are
identical.
8.1.2 Molecular Orbitals of CH4
CH4 belongs to the tetrahedral symmetry Group Td ; we may put the C atom
at the origin and H atoms at (-1,1,1),(1,-1,1),(-1,-1,-1),(1,1,-1) in appropriate
units. There are NG =24 operations. There are C3 and C32 rotations around
159
4
3
Fig. 8.2. Tetrahedral CH4 molecule inscribed in a cube; the open circle denotes
the C position.
Td
A1
A2
E
T1
T2
E
1
1
2
3
3
8C3
1
1
1
0
0
3C2
1
1
2
1
1
6d
1
1
0
1
1
6S4
g = 24
1
r2
1
0 (3z 2 r2 , x2 y 2 )
1
(Rx , Ry , Rz )
1
(x, y, z)
For the Hydrogens, let d(i, C) denote the list of the destinations of atom 1
under the operations in class C. Here is the situation:
160
Table 8.1.2 and the Character Table allow to project on T2 by the operator
(7.70):
P (T2 ) |1 = 3|1 (|2 + (|3 + |4 ) + [3|1 + (|2 + |3 + |4 )] 2(|2 + (|3 + |4 )
= 2[3|1 (|2 + (|3 + |4 )].
If we wish to obtain z the fastest way is probably to symmetrize between
1 and 2. Indeed, if 3|1 (|2 + (|3 + |4 ) is in T2 , 3|2 (|1 + (|3 + |4 )
also is; we sum and normalize obtaining
|z =
Also,
|1 + |2 |3 |4
.
2
|1 |2 |3 + |4
,
2
|1 |2 + |3 |4
|y =
.
2
|x =
J
MJ =J
exp(iMJ ) =
sin((J + 12 ))
.
sin( 2 )
(8.5)
All rotations by an angle belong to the same class, thus the character is
this, independently of the rotation axis.
8.1.4 Examples: Oh Group, Ligand Group Orbitals, Crystal Field
The Oh Group
We consider a cube with a system of cartesian axes through the centers of the
faces. The centers of the faces are vertices of an octahedron. The operations
of symmetry of the cube and octahedron are the same. We label the vertices
of the octahedron as in Figure 8.1.4. There are 48 operations in the cubic
Group Oh . The C4 rotations around the cartesian axes and their inverses
form the 6C4 class, and the C2 rotations form the 3C2 class. The 2
3 rotations
in both senses around the 4 cube diagonals give a class 8C3 ; these axes are
perpendicular to faces of the octahedron. Joining the center of EC with the
center of FB we have a C2 axis. There are 12 edges, so there are 6 axes and the
class is 6C2 . The inversion forms a class; the 3h class includes reections like
161
the one in the BCEF plane. The C4 rotations followed by a h reection in the
plane perpendicular to the rotation axis form the 6S4 improper rotation class.
The plane containing A,D and the midpoints of EC and BF is a symmetry
plane, and the reection belongs to a 6d class. Finally,there are S6 improper
rotations: the axes are those of the C3 rotations; if we look along a C3 axis,
opposite faces appear as concentric triangles; the gure is invariant under
a C6 rotation followed by a reection in the orthogonal plane through the
center (Figure 6.2). With 4 axes and two senses of rotations one builds a 8S6
class.
z
A
C
y
D
Fig. 8.3. Octahedron
6 2
3C2
1
1
1
1
2
2
1
1
1
1
2
6C2
1
1
1
1
0
0
1
1
1
1
0
8C3
1
1
1
1
1
1
0
0
0
0
0
i 6S4
1 1
1 1
1 1
1 1
2 0
2 0
3 1
3 1
3 1
3 1
0 0
3h
1
1
1
1
2
2
1
1
1
1
4
6d
1
1
1
1
0
0
1
1
1
1
2
8S6
g = 48
1
x2 + y 2 + z 2
1
1
1
1 (x2 y 2 , 2z 2 x2 + y 2 )
1
0
(Rx , Ry , Rz )
0
(x, y, z)
0
(xy, xz, yz)
0
0
The Oh Group is the point Group of many interesting solids, including complexes like CuSO4 5H2 O and FeCl3 where a transition metal ion at the center
162
AD
.
2
CF
BE
Crystal Field
The number of d electrons in transition metal ions is:
Ti V Cr Mn Fe Co Ni Cu
M ++ 2 3 4 5 6 7 8 9
M +++ 1 2 3 4 5 6 7 8
Hunds rule allows to nd the ground state of partially lled d shells and
the corresponding number of unpaired electrons as follows:
d1 d2 d3 d4 d5 d6 d7 d8 d9
ground state D 3 F 4 F 5 D 6 S 5 D 4 F 3 F 2 D
unpaired
1 2 3 4 5 4 3 2 1
2
for isolated transition ions one would always predict paramagnetism, but the
compounds can be paramagnetic or not. The bivalent Fe ( 3d6 conguration)
163
forms the complexes [F e(H2 O)6 ]2+ , which is green, and paramagnetic, but
also [F e(CN )6 ]4 , which is yellow and diamagnetic. These facts can already
been understood by the crystal eld theory, in which the ligands behave like
point charges, or anyhow they generate a eld to octahedral symmetry that
resolves the degeneracy of the atomic terms (for the rst transition series the
spin-orbit interaction is a small perturbation to be introduced subsequently).
A more quantitative treatment is obtained then from the ligand eld theory.
For d (or D) states, from (8.5) and from the even parity we nd the
characters
Oh E 6C4 3C2 6C2 8C3 i 6S4 3h 6d 8S6
d 5 1 1 1 1 5 1 1 1 1
and conclude d = Eg T2g . This is already clear from the Character Table,
reporting (dxy , dxz , dyz ) as a basis for T2g and (dx2 y2 , z 2 ) for Eg . In transition
metal complexes usually = E(Eg ) E(T2g ) > 0, since the T2g orbitals stay
far from the negative ligands. Thus in the absence of Coulomb interactions
one would ll the available levels according to the aufbau principle, starting
with T2g . In crystal eld theory one tries to predict the magnetic properties by
diagonalizing a many-electron Hamiltonian which is the sum of the isolated
ion Hamiltonian and the one of the crystal eld.
There are two simple limiting cases. If
U, where U represent the
order of magnitude of the multiplet splitting, due mainly to the Coulomb
interaction, one treats as a perturbation of the isolated ion multiplet. The
atomic Hund rule holds, and paramagnetism obtains. If U, Hunds rule
holds (high spin is preferred) within the degenerate T2g and Eg levels, but Eg
starts being lled only after T2g is full, and 6 electrons yield a diamagnetic
complex.
(8.6)
where I introduced a notation vi for the generic cartesian coordinate. All the
possible motions of the nuclei are described classically from the equations of
Newton,
164
mi vi =
U
vi
(8.7)
where mi is the mass associated to v in the obvious way x1, y1, z1 m1 and
so on. In the harmonic approximation, expanding around to one conguration
of equlibrium,
U=
1
Uij vi vj ,
2 ij
(8.8)
1
Uij vj .
mi j
(8.9)
Qi = vi mi
(8.10)
Uij
Wij =
mi mj
(8.11)
(8.12)
The eigenvectors Q of the W matrix are the normal modes, and their eigenfrequencies are obtained from the secular equation. Three modes, of null
frequency, correspond to rigid translations of the molecule, and 3 others (or
2, for linear molecules) to rigid rotations. Rotations also have = 0, since
the energy of the molecule does not depend on its orientation. The remaining
3N-6 (or 3N-5) frequencies are vibrational. The Group theory is helpful to
simplify the solution of the secular equation. Every R operation of the Group
of the molecule sends every nucleus in itself or to another identical nucleus in
an equivalent position; meantime, it produces a (proper or improper) rotation the system of cartesian axes that we can imagine xed at every nucleus.
So, R maps each Qi into a linear combination of the components. In such
a way, we associate to every R a matrix D(R), and obtain a representation
of the Group in the space of the vectors Q. Every D(R) commutes with
W , since R is a symmetry and W must be invariant under R; thus if Qa
is a solution of the secular equation, also D(R)Qa must be a solution, and
with the same . Therefore, we can diagonalize W simultaneously with the
maximum number of commuting D(R) matrices. The theory of the Groups
adds the Dirac characters; this means that in a new basis whose elements
are vibrations pertaining to irreps of the Group, W is block diagonal. Thus,
the normal modes can be assigned to irreps of the Group; the reduction of
165
0 0 0 0 0 0 1 0 0
0 0 0 0 0 0 0 1 0
0 0 0 0 0 0 0 0 1
0 0 0 1 0 0 0 0 0
(8.13)
D(C2) =
0 0 0 0 1 0 0 0 0 .
0 0 0 0 0 1 0 0 0
1 0 0 0 0 0 0 0 0
0 1 0 0 0 0 0 0 0
0 0 1 0 0 0 0 0 0
D(C2 ) and D(v (yz)) have the same block structure
00b
D(C2 ) = 0 b 0
b00
where now 0 is for a null 3 3 block, and
1 0 0
1 0
b(C2 ) = 0 1 0 , b((yz)) = 0 1
0 0 1
0 0
(8.14)
0
0;
1
(8.15)
(8.16)
We could arrive to this result without writing the D matrices, taking into
account that for each operation:
166
C2v
A1
A2
B1
B2
I
1
1
1
1
C2 v v
1 1 1
1 1 1
1 1 1
1 1 1
g=4
z
xy, Rz
x, Ry
y, Rx
trasl
rot
vibr
E
9
3
3
3
C2 (xz) (yz)
-1 3
1
-1 1
1
-1 -1
-1
1
3
1
(8.18)
167
vibr = 2A1 B1 .
Finally. The water molecule has two total-symmetric vibrations and one of
B1 symmetry.
A total-symmetric vibration, that all the molecules have, is the breathing
mode. In the symmetric stretch the angle between the OH bonds varies. In
order to discover what kind of vibration is the one labeled B1 that goes like
x one can construct the projection operator
P (B1 ) = E C2 + (xz) (yz);
in terms of block representative matrices,
1 + b((xz))
0 b(C2 ) b((yz))
,
0
0
P (B1 ) =
1 + b((xz))
b(C2 ) b((yz)) 0
(8.19)
(8.20)
I
1
1
2
2C3
1
1
1
3v g = 6
1
z
1 Rz
0 (x, y)
The Cartesian movements of N are transformed like the coordinates. Therefore for N we have:
C3v E 2C3 3v
(8.21)
(N ) 3 0 1
Lets nd the characters of (H3 ). (E) = 9 (no arrows move) (C3 ) = 0
(all the H move and contribute 0) (v ) = 1 (two H reected one in the
other: character 0; for the other H, two arrows in the reection plane and one
orthogonal: character 1) Therefore:
C3v
E 2C3
(N ) 3 0
(H3 ) 9 0
(N H3 ) 12 0
trasl 3 0
rot
3 0
vibr 6 0
Finally, vibr = 2A1 + 2E.
3v
1
1
2
1
-1
2
168
E 8C3
1 1
1 1
2 1
3 0
3 0
15 0
3 0
3 0
9 0
3C2
1
1
2
1
1
1
1
1
1
6d
1
1
0
1
1
3
1
1
3
6S4
g = 24
1
r2
1
0 (3z 2 r2 , x2 y 2 )
1
(Rx , Ry , Rz )
1
(x, y, z)
1
1
= T2
1
= T1
1
A1 E 2T2
D6h
A1g
A2g
B1g
B2g
E1g
E2g
A1u
A2u
B1u
B2u
E1u
E2u
tot
trasl
rot
vibr
E 2C6
1 1
1 1
1 1
1 1
2 1
2 1
1 1
1 1
1 1
1 1
2 1
2 1
36 0
3 2
3 2
30 4
2C3
1
1
1
1
1
1
1
1
1
1
1
1
0
0
0
0
C2 3C2
1 1
1 1
1 1
1 1
2 0
2 0
1 1
1 1
1 1
1 1
2 0
2 0
0 4
1 1
1 1
2 2
3C2
1
1
1
1
0
0
1
1
1
1
0
0
0
1
1
2
i 2S3
1 1
1 1
1 1
1 1
2 1
2 1
1 1
1 1
1 1
1 1
2 1
2 1
0 0
3 2
3 2
0 0
2S6
1
1
1
1
1
1
1
1
1
1
1
1
0
0
0
0
h 3d
1 1
1 1
1 1
1 1
2 0
2 0
1 1
1 1
1 1
1 1
2 0
2 0
12 4
1 1
1 1
12 4
169
3v
g = 24
1
1
Rz
1
1
0 ! (Rx , Ry ) "
x2 y 2 , xy
0
1
1
z
1
1
0
(x, y)
0
0
1
1
0
The Group has 24 elements, and it is found thatvibr = 2A1g A2g A2u
2B1u 2B2g 2B2u E1g 3E1u 4E2g 2E2u .
k =
N
with p, q, r Z, and the reciprocal lattice vectors dened by ti gj = 2ij .
We know from Sect. 7.2.1 that Blochs functions are
x ),
(8.22)
x ) = ei k x u
(
(
k
k ,
where u
x ) are lattice periodic. Moreover, since (letting h = 1)
(
k
p ei k x = ei k x (
p + k ),
we may write
170
(
p + k )2
x ).
+ V ( x ) u
x ) = k u
(
(
k
k
2m
(8.23)
+ V (
x ) + HSO
(8.26)
H=
2m
with
1
V
p.
(8.27)
HSO
=
4m2 c2
Eigen-spinors can be taken of the Bloch form,
x)
(
x
x
i k
i k
k ,+
u
(8.28)
(x) =e
(x)= e
u
k ,
k ,
(x)
k ,
One nds spinors with = + and - that reduce to up and down spin for
c , but otherwise spin and orbital degrees of freedom mix. The periodic
functions are obtained by solving
1
(
p + k )2
+V(x)+
V ( p + k ) u
x)
(
k ,
2m
4m2 c2
x ). (8.29)
=
(
u
k , k ,
Space Inversion
Schr
odinger equation with a periodic potential V ( x ); an inverted crystal
x ) = P 0 V (
x ) = V (
x ) instead. Since
poses the same problem with V (i) (
(P 0 )2 is the identity, from Hk (x) = k k (x) one gets
P 0 HP 0 P 0 k = H(x)P 0 k = k P 0 k ;
171
x ), one cannot
is the same as in the original problem. If V ( x )
= V (
k
speak of degeneracy, since the corresponding problems remain distinct. If
however the crystal has inversion symmetry, that is, [P 0 , H] = 0, adding
this element to the translations produced a non-Abelian Group which implies
degeneracy. Indeed, k (x) and k (x) both belong to the eigenvalue
.
k
0
ikx
However, P k (x) k (x) = e
uk (x) belongs to k since it gets
multiplied by eikt under a translation by t. Therefore we must make the
identications
(8.30)
P 0 k (x) = k (x),
and uk (x) = uk (x); so,
=
.
k
k
(8.31)
Time Reversal
The antilinear1 Kramers operator K takes the complex conjugate, and is
useful to discuss time reversal symmetry, as I show in a moment. In the
Schr
odinger theory, assume one knows how to solve a time-dependent problem
i
h
(t)
= H(t)(t)
t
(8.32)
(t )
= H(t ) (t ).
t
(8.33)
Things are more involved with the Pauli equation; I write H0 the spinindependent part of the Hamiltonian, which I assume real, and separate the
spin-eld coupling:
i
h
1
eh
(t)
= [H0 (t) + B(t)](t), =
.
t
2mc
is antilinear if O(a
+ b O).
An operator O
+ b) = a O
(8.35)
172
Taking the complex conjugate and multiplying both sides by i2 =
0 1
1 0
one nds
Note that
K, K = , ,
(8.39)
T , T = , .
(8.40)
and
Then
T 2 = ,
thus
T 1 = T .
Moreover T ; indeed, if =
,
T = iy K
173
0 1
=
=
1 0
T | = (, )
= 0.
then
T
x ) = iei k x y u
x)
(8.41)
(
(
k
k ,
have the same energy and belong to k , ( they are periodic spinors times
x ) = T z
x ), T
x ) ;
x ), z
(
(
(
(
k
k
k
k
the anticommutation rules give
x ), T
x ) ;
= z T
(
(
k
k
since z is Hermitean,
x ), z
x ) = T
x ), z T
x ) .
(
(
(
(
k
k
k
k
In summary the time reversal invariance requires
x) =
x );
T
(
.
=
(
k
k ,
k ,
k ,
(8.42)
Conjugation
If P 0 and T are simultaneous symmetries,the conjugation
C = P 0T
(8.43)
also is conserved. Using equations (8.30),(8.42) we see that this is such that
C
x ) =
x );
(
(
=
k ,
k ,
k ,
k ,
(8.44)
r +
a
(|a) :
r =
(8.45)
174
(| b )(|
a ) = (| b +
a)
(8.46)
and it has the Point Group as a subgroup. We can write the faithful representation2
a
r
r
=
.
(8.47)
1
0 1
1
The inverse operation is
Since
(|
a )1 = (1 | 1
a ).
(8.48)
a ) = (1 |1 ( b +
a
a ))
(|
a )1 (| b )(|
(8.49)
r +
a;
r =
(8.50)
s b +
a + b.
s =
(8.51)
a.
b = ( 1)1
(8.52)
175
s , if 0 = (1 ) b +
a.
the transformation becomes homogeneous,
s =
(|
a )n = (n |
n1
a ) = (1, t ).
k
(8.53)
k=0
a where
a is normal to the rotation axis. The
Let us write
a =
a +
.
a =
n
(8.54)
176
41 2
3
d m
is not; d denotes a glide plane with a translation 1/4 Bravais vector.
F
r ); so
ei k t . In general, one denes Rf (
r ) = f (R1
(,
a )ei k r = ei k (, a ) r
and using the inverse operation
(,
a )ei k r = ei k ( r a ) .
Rotating two vectors by the same angle the scalar product does not change;
so we may write
(,
a )ei k r = ei( k )( r a ) .
(8.55)
stars with fewer elements. For non-degenerate bands, k points of low symmetry correspond to a single , but in general there is a subspace associated
to a given k point.
k = k + G
G
has the invariant Subgroup T
T G of the translations t such that
k
k
177
(8.57)
A(2, 3)S(1, 3) and A(2, 3)S(1, 2) are mixed symmetries and project on E.
These projectors written in terms of symmetrizers and antisymmetrizers are
called Young projectors. They may be thought of as projectors on the various
irreps based on the regular representation. The three diagrams exhaust the
possible partitions of 3 in not increasing integers, that is, 3 =2+1=1+1+1.
The Young tables or Young tableaux are obtained from the Young diagrams by inserting numbers from 1 to 3 so that every line and every column
grow along. The tableau
123
178
represents the projection operator S(1, 2, 3), the symmetrizer ; the Young
12
but one can also
projection operator A(1, 3)S(12) is given by
3
1
13
antisymmetrize with respect to 2, getting
. Finally, 2
2
3
projects on A2 .
The fact that there are two tables with mixed permutation symmetry is
due to degeneracy 2 of the irrep E. In general, in the Young tables for S(N ),
the m-dimensional irreps occur m times.
Young Tableaux for S(4)
S(4), has 24 elements and the following 5 irreps that may be found by the
above stated rules.
1234
123
4
134
2
13
24
is 2-dimensional,
1
2
is 3d and nally
is the totally antisymmetric ir3
4
rep. These correspondences are useful as it extends to all groups S(N) of
permutations of N objects. Thus, the irreps of S(N) are known for any N.
12
3
4
13
2
4
14
2
3
179
Problems
3
2
5
a)
b)
Fig. 8.4. a) The CuO4 Hubbard model cluster. b) The Cu5 O4 Hubbard model
cluster.
(9.2)
182
T(1),(2),,(r)
=
(1),(2),,(r)
(9.3)
Given a second-rank tensor T , one can build a symmetric tensor Sij =
Tij + Tji and an antisymmetric one Aij = Tij Tji . Obviously a symmetric
(antisymmetric) tensor remains symmetric (antisymmetric) under the transformations of GL(n). In general, the tensors of GL(n) are reduced into irreducible parts by taking linear combinations according to the irreps of the
permutation Group S(r). Further reduction is possible in subgroups of GL(n).
Tensors in Polar Form
Under the operators of O+ (3) the ITO transform like spherical harmonics.
Therefore, the following polar form is expedient. For a vector, the polar components are xm , m = 0, 1, where, using a traditional notation,
x + iy
x iy
x+1 = , x0 = z, x1 = ,
2
2
and transform according to
Rxm R1 =
1
xn Dnm
(R),
(9.4)
(9.5)
n
1
(R) represents the rotation in the basis of the spherical harmonics
where Dnm
with l = 1. It is natural to put the basis vectors in polar form too,
e1 + ie2
e1 ie2
, e0 = e3 , e1 =
.
(9.6)
2
2
3
In this way the expansion of a vector in the basis, V = i=1 Vi ei is replaced
by
1
()m Vm em .
(9.7)
V =
e+1 =
m=1
1
()m Vm Wm ,
(9.8)
m=1
as one can readily verify. The advantage of the polar form is that now we can
generalize to an ITO of any rank of O+ (3). The denition of such an ITO is
()
Tq() Dqp
(R).
(9.9)
RTp() R1 =
q
183
a12 b12
a12 b22
.
a22 b12
a22 b22
R|ij =
n
m
p
k
n
m
k
()
()
(9.11)
Thus, the direct product matrix D() = D() D() with elements
()
()
= Dki (R)Dpj (R) is the representative matrix of R in what is
called the direct product representation. Its characters are
()
Dkp,ij (R)
184
( (R) =
()
(9.12)
kp
I
1
1
1
1
2
1
1
2
4
C2 2C4
1 1
1 1
1 1
1 1
2 0
1 1
1 1
2 0
4 0
2 v
1
1
1
1
0
1
1
0
0
2d
g=8
1
z
1
Rz
1
x2 y 2
1
xy
0
(x, y)
1
B2
1
A2
0
E
0 A1 A2 B1 B2
(9.13)
In this way we can build a multiplication table for the irreps of C4v .
C4v
A1
A2
B1
B2
E
A1
A1
A2
B1
B2
A2
A2
A1
B2
B1
B1
B1
B2
A1
A2
B2
B2
B1
A2
A1
E
E
E
E
E
A1 + A2
E E E E
+B1 + B2
(9.14)
1
1 () ()
1 () (R) =
(R) = .
NG
NG
R
(9.15)
185
where Rpq xp xq is the Raman tensor. From the symmetry viewpoint, what
matters is that the components of the Raman tensor transform like xp xq ,
and this determines the selection rules. In systems with inversion symmetry,
the normal modes must be gerade or ungerade (even or odd). Only ungerade
modes are infrared active and gerade ones are Raman active. For instance,
infrared and Raman spectra of Benzene have no frequencies in common.
186
inserting a complete set one can go, with a unitary transformation, from the
basis {|ij } of the direct product () () to a basis of functions that
transform according to irrep () of the Group G.
|ij =
m
irreps
g
|r r|ij .
(9.16)
()
()
(9.17)
rs
This corresponds to the familiar use of CG coecients for the sum of angular
momenta (for an example, see Problem 9.1).
e 2 0
i
z
2
(9.18)
R = e
=
= cos( ) iz sin( )
0 ei 2
2
2
and belongs to the SU(2) covering group of SO(3). For = 2, R = 1. If
we rotate around axis n, since ( n)2 = 1,
R = ei 2 n = cos( ) + i( n) sin( )
(9.19)
2
2
and for = 2, R = 1 anyway2; this is a rotation that commutes with any
other symmetry.
2
187
ei 2 0
(1/2) () = 2 cos( ).
(9.20)
2
If the inversion i is in G, it leaves spin and any angular momentum invariant,
so D(i) = D(E) and (i) = 2. The reections and all the improper rotations can be written like products iR .3 In such a way, we can complete the
characters of the spinorial representation4. This describes an electron with
an orbital A1.
The above information is enough to build easily the character table for
G from that of G, without having to work out everything from the multiplication table. Having listed the classes, one can append the irreps of G,
C3v
E E
A1
A2
E
E1/2
5
6
3
1
1
2
2
1
1
1
1
2
2
1
1
C3 C32
C3 E 3v 3v E
C32 E
1
1
1
1
1
1 1 1
1 1 0
0
1
1 0
0
1
1
i
i
1
1 i
i
188
The A1 , A2 and E irreps have the same charcters as in C3v , and E1/2 is
the spinor representation. Since the classes are 6 two irreps are missing, and
since the sum of the squares of the dimensions must be 12, the new ones
are one-dimensional. By orthogonality, we nd that they are conjugate representations as shown above. At this point the reader could solve Problem
9.4.
In a physical problem, one can begin by classifying the space orbitals
according to G and then extend the theory to G including spin. The direct
product of the orbital irrep by the spinor representation will include the
spin. In general , the result will be reducible in the double Group applying
the LOT. We will be able to thus establish how the spin-orbit interaction
reduces the degeneracy in the problem in issue. For example, if the orbital
belongs to A1 or A2 of C3v , the spinor belongs to E1/2 , and no level splitting
occurs; if the orbital belongs to E the product representation has dimension
, E E1/2 = E1/2 + 5 + 6 .
4, but one nds that in C3v
(9.21)
where Te (r) and TN (R) are the electronic and nuclear kinetic terms, and
V (r, R) = Vee + VeN + VN N
(9.22)
(9.23)
(9.24)
(9.25)
yields the adiabatic eigenstates n (r; R) and the potential energy surfaces
En (R). This is the BO approximation, which further assumes that if nuclei
move their evolution is conned to an adiabatic surface, and the harmonic
oscillations about equilibrium correspond to the minimum of E0 (R).
189
2 2
h
+ E0 (R)(R) = W (R).
2M R2
(9.26)
the positions of the nuclei are external parameters that determine the
Coulomb external potential in which the electrons move, and the Hamiltonian H(R)
the momenta i
hR canonically conjugated to the nuclear positions are
ignored. This assumes that the nuclear masses are innitely large
The equilibrium conguration corresponds to a minimum of the total
energy:
(9.27)
E0 (R) = 0 |H|0
E0
= 0.
R
(9.28)
0
( E
R = 0 represents extremum conditions in all the components of the
nuclear position vectors)
the electronic states are calculated at the equilibrium conguration and
E0 (R) is the potential energy for the nuclear motion.
190
The CH4 ground state is non-degenerate in the Td symmetry; but by removing a bonding t2 electron the CH4+ ion distorts until the symmetry Group
becomes Abelian.
Initially proposed as a computational aid which exploits a given symmetry
of the Hamiltonian, Group Theory eventually dictates which symmetries are
allowed or forbidden at all.
Mathematical Formulation of the Problem
of the nuclei in
We want a criterion to decide if a given conguration R
the BO approximation can be the equilibrium one, according to (9.28). We
must minimize with respect to the shifts of the nuclei from to the reference
but rst, we eliminate the rigid shifts of the molecule by
conguration R;
using the normal modes of vibration of Section (8.2) instead of the nuclear
positions. These modes are labeled by the index of the frequency and by
an index i for the degeneracy and multiplicity (the same symmetry can occur
several times). The amplitude of the motion according to a normal mode
is specied by a normal coordinate qi , and we must minimize E(q), where
q {qi } stands for the whole set.
Actually, we do not have any analytic expression of E(q) to dierentiate.
in powers of q, letting
Therefore, we expand the Hamiltonian H around R
ik
Since the qi
(9.31)
6
Actually, they could in principle only increase it: with a A1 motion the water
molecule could be straightened and become a Dh molecule.
191
Non-degenerate Case
If the ground state is not degenerate, (0 0 ) = A1 and 0 |Vi |0 = 0 for
= A1 . So, all the nuclei move, keeping the symmetry of the molecule, until
(A )
0 |Vi 1 |0 = 0. This 0 is not due to the symmetry, but to the existence of
a minimum of energy versus breathing mode coordinate: we can vary a bond
length or an angle until the condition is satised.
Degenerate Case
(0)
(0)
(0)
When H0 = E0 is solved by several , we must apply degenerate perturbation theory, diagonalizing the perturbation matrix with elements
(0)
(0)
|H | . If the eigenvalues do not vanish all identically, one predicts
()
a splitting linear in the qi of the degenerate level, a lowering of the symmetry and a distortion of the molecule. The only Hermitean matrix that has
all eigenvalues equal to 0 is the null one. Thus, for equilibrium we need to
satisfy the strong condition
(0) |Vi |(0) = 0 , , .
(9.32)
Now we must examine the normal modes occurring in the assumed geometry
in order to see if any generates matrix elements that destroy the symmetry.
(0)
Let 0 denote the irrep where belongs. The Jahn-Teller eect is caused
by a normal mode belonging to an irrep
= A1 which is contained in 0 0 .
In such cases there is no reason why the matrix elements vanish, and the
molecular conguration is unstable7 .
In 1937, Jahn and Teller demonstrated that for a non-linear molecule
with degenerate ground state irrep 0 , a vibration always exists such that
= A1 is contained in 0 0 ; this implies that the fundamental electronic
terms of non-linear molecules are not degenerate (even if not necessarily totalsymmetric). The proof is obtained by repeating for all the 32 point Groups8
the same analysis that we now exemplify in the case of Td .
Example
We resume CH4 (Sections 8.1.2,8.2);
Td
A1
A2
E
T1
T1
7
E
1
1
2
3
3
8C3
1
1
1
0
0
3C2
1
1
2
1
1
6d
1
1
0
1
1
6S4
g = 24
1
r2
1
0 (3z 2 r2 , x2 y 2 )
1
(Rx , Ry , Rz )
1
(x, y, z)
192
(9.33)
E
1
1
1
1
2
C2 2C2
1 1
1 1
1 1
1 1
2 0
2d
1
1
1
1
0
2S4
1
1
1
1
0
193
(9.36)
h2
2
+
dR (R)E0 (R)(R)
F [] =
drdR 0
0
2M
R2
(9.37)
W dR (R)(R)
where W is a Lagrange multiplier that ensures normalization. Using
2
2
2
|0 |
+
|
|
=
|
|
+
2
|0
0
0
R2
R2
R R
R2
9
(9.34)
with
He (r; R0 )n (r; R0 ) = E(R0 )n (r; R0 ),
the topology of the surfaces is important. For conical intersections, when the surfaces cross each other, one speaks of Jahn-Teller eect; when surfaces touch at extremal points one speaks of Renner-Teller eect [5]. However, the JT theorem limits
the occurrence of glancing intersections to linear molecules (and to cases when the
gradient is accidentally vanishing or particularly small for reasons independent of
Group theory).
194
one nds
F [] =
+
+
2
h
2M
dR
dR
2
+2
R2
0
dr0
R2
drdR 0
0
R R
(9.38)
dR (R)E0 (R)(R) W
dR (R)(R).
We vary the bra (variation of bra and ket produces identical results).
h2 2
h2
0
F = dR (R)
dr0
2
2M R
M
R R
2
2
h
(R)
W
)(R)
= 0. (9.39)
(R)
+
(E
dr0
0
2M
R2
This implies:
h2
2 2
h
0
+ E0 (R)(R)
dr0
2M R2
M
R R
2
h2
(R) = W (R)
dr0
2M
R2
(9.40)
h2
0
0
(R) =
.
dr0
dr0
M
R R 2M
R2
(9.41)
(9.42)
This has been often ignored in the literature; the reasons are that 1) the rst
contribution, averaged over real electronic wave functions vanishes since
1
0 (r; R) =
dr0 (r; R)
dr0 (r; R)0 (r; R);
R
2 R
m
). It will be apparent shortly
2) the second contribution is small (of order M
that such reasons are not generally as safe as they may appear to be.
195
Then, the kinetic energy of the nuclei does not play a role, and one can
observe a static JT eect with broken symmetry. At weak coupling, one
speaks about dynamic JT eect and the system oscillates between several
minima; the overall symmetry remains unbroken10. The JT approximation
to the solutions of Htot tot = W tot ( Equation 9.23) uses as a reduced basis
set the Ndeg degererate adiabatic functions in the symmetric conguration
R0 :
Ndeg
n (R)n (r)
(9.43)
tot (r, R)
n
here n (r) are assumed known, and one seeks the nuclear amplitudes n (R).
Substituting into Equation (9.23) and taking the scalar product by m | leads
to
(r; R0 ) [Htot W ]
n (R)n (r) = 0,
(9.44)
drm
that is,
drm
n (R)n (r) = 0,
TN + Te (r) + V (r, R) W
:
;<
=
2 2
h
m (R) +
n (R)
2
2M R
n
drm
(r; R0 )He (r, R)n (r) = W m (R).
(9.45)
Dening
Vmn (R) =
drm
(r; R0 )He (r, R)n (r)
(9.46)
2 2
h
(R)
+
Vmn (R)n (R) = W m (R).
m
2M R2
n
(9.47)
or in matrix form
10
The distinction between dynamical and static JT eect usually depends only
on the time scale of the experiment. For example, (Cu6H2 O)++ ions look perfectly
octahedral when observed at room temperatures in EPR experiments, but below
20 0 K or with fast spectroscopies it is seen that this symmetrical conguration is
the time average of stretched and compressed ones as the top and bottom H2 O
molecules oscillate up and down. On the other hand, X-Ray diraction at room
temperature shows that (CuBr6 )4 ions is a tetragonally distorted, stretched octahedron; the latter is classied as an example of static JT eect. It should be kept
in mind, however, that the JT approximation may fail completely, as it does in
strongly correlated models with strong electron-phonon coupling [71].
196
V11 (R) .
.
.
=
+
.
.
2M R2
V1 (R) .
2
HJT
.
.
.
.
. V1 (R)
.
.
.
.
.
. V (R)
(9.48)
(9.49)
Here, V (x), V (y) transform according to E and represent the operator poten must belong to A1 and
tial due to the phonons acting on the electrons; K
2
2
2
q = qx + qy . Then, (9.48) becomes
HJT =
2 2
h
h2 2
(V (x))xx (V (x))xy
+
q
x
(V (x))yx (V (x))yy
2M qx2
2M qy2
(V (y))xx (V (y))xy
2.
+qy
+ K q
(V (y))yx (V (y))yy
(9.50)
Group theory dictates the form of the vibronic interaction; for illustration, we
adopt the geometry of Figure 7.1 of Section 7.2.2. (x, y) is a basis for E; for
the present purpose, however, we shall use an alternative basis for E in C3v ,
with the same D matrices, namely, (fx , fy ) = (2xy, x2 y 2 ). It is is evident
that 2xy transforms like x and x2 y 2 like y in the chosen geometry, since
2xy is odd and x2 y 2 even under the 1 reection. For any basis (fx , fy ) of
E, we know from (7.46) that fx |y = fy |x = 0 and that fx |x = fy |y .
Therefore, if now (x , y ) are electronic states and are a basis for E, x2 + y2
belongs to A1 and for the V (x) elements we nd:
2 drx y V (x),
dr(x2 y2 )V (y) = 2 =
2
2
dr(x2 + y2 )V (y) = 0 = dr(
(9.51)
x + y )V (x),
2
2
dr(x y )V (x) = 0 = 2 drx y V (y).
Thus, V (x) has equal o-diagonal matrix elements on (x , y ) and nothing
on the diagonal and V (y) has opposite diagonal elements and 0 o-diagonal;
this is clear already when one considers the a parity. Thus, the form of the
JT hamiltonian in the E problem is:
HJT =
11
2 2
h
h2 2
x2 + qy2 ).
+ [qx x + qy z ] + K (q
2M qx2
2M qy2
(9.52)
197
with
M=
sin() cos()
cos() sin()
(9.54)
.
198
RTp() R1 =
()
Tq() Dqp
(R).
(9.57)
As the components i,p and k vary, one nds a number of matrix elements
()
< i|Tp |k > that are all connected by the Wigner-Eckart theorem
Theorem 9.
(9.58)
where the reduced matrix element < T () > does not depend on the
components;i|pk is the Clebsh-Gordan coecient.
Since the Clebsh-Gordan coecients are mere geometry, the dynamics
enters through the reduced matrix element. A formal proof is given in Appendix C, but an intuitive argument is also useful. A very crucial point in
that the tensor and the function spaces must be irreducible. If you know one
basis function of an irrep you can build all of them by the Group operations
and orthogonalization; and the same is true of the tensor components; then
it is at least very plausible that all the matrix elements can in principle be
obtained by symmetry from the knowledge of any non-vanishing one of them.
Based on this, the theorem is then most simply understood. Suppose that by
()
direct computation we obtain a particular element i0 |Tp0 |k0 = Q
= 0.
Next, we decide to look at the system from a new transformed reference ob()
tained by some R G; then, |i0 , |k0 and Tp0 transform to linear
combinations of all the components, still remaining in their irreps; however
the matrix element is still Q. Varying R, we can write a system of linear
equations linking the components, all with the same right-hand-side. Are all
those equations must be compatible, and since there are enough relations to
()
determine < i|Tq |k >, one can solve and each matrix element must be
proportional to the only r.h.s. Q. Any two tensors T and T of the same irrep generate the same system, except that the r.h.s. are Q and Q; so they
must yield proportional results. There is a particularly simple tensor dened
by Tp |k = |pk which yields the theorem with < T () > 1. For all
the other tensors the theorem holds with some reduced matrix element.
Simple Applications
The theorem reduces the calculation of a tensor to one of one its components.
It implies less that in a symmetry adapted basis all the irreducible tensor
operators have the same elements matrix elements, up to a multiplicative
constant. We can choose the most comfortable operator (as long as it does
not have a null reduced matrix element). As an example, in O(3), j = 1 labels
(1)
the irrep of vectors and Tq , q = 0, 1 is a vector operator in polar form.
Therefore, its matrix elements are proportional them to those of j, that they
are easy to calculate in this basis:
jm|T |j m = Cjm|J |j m .
199
(9.59)
jm|J|jm jm |J|jm = CJ(J + 1).
(9.61)
m
(9.62)
3Q
2
Ii Ik + Ik Ii ik I 2 ,
2I(2I 1)
3
(9.63)
built with a constant Q (named the quadrupole moment) and withthe components of the nuclear spin I. In a similar way one can build an atomic
quadrupole moment tensor from J .
The spin-orbit interaction is HSO = i l(i)s(i) in many-electron atoms,
where l(i) is the orbital angular momentum of electron i , s(i) is its spin;.
Since l(i) is a vector like the total orbital angular momentum L and s(i) is
a vector in spin space like the total spin angular momentum S, the WignerEckart theorem allows to write
LM LS MS |HSO |LM LS MS = ALM LS MS |L S|LM LS MS ,
(9.64)
where A is a constant.
Pi
m
j
()
()
j Dj,i (P).
(9.65)
200
m
()
()
n Dn,q (P).
(9.66)
When permuting the electrons, i.e. doing the same permutation on spins and
coordinates, the wave function must go into itself up to a sign, therefore
components must enter symmetrically, like in
=
m
() ()
k k .
(9.67)
m
()
()
[Pk ][Pk ];
(9.68)
()
()
(9.69)
()
(9.70)
Two irreps that satisfy the condition (9.70) are conjugate or associate. We
saw in Sect. 8.5 that the spin eigenfunction have a Young tableau consisting
of up to two lines. The following [N M, M ] two-line tableau (left) with
N M integers denoted ai aN M on the top line and M integers denoted
bi bM on the bottom line is suitable for a spin eigenfunction symmetry of
S = 126(N 2M ) (Equation 9.71, left); then, the conjugate tableau is
5spin
M
2 , 1N M as shown below (right).
a1 a2 . . . . . . . aN M
b1 b2 . . . . . bM
a1
a2
.
.
.
.
.
.
.
.
aN M
b1
b2
.
.
.
.
.
bM
(9.71)
201
(9.74)
m
n
= Tmn Tnm
(9.75)
1
1
GM=4 = (2, 2)
GM=3 = (1,2)+(2,1)
2
GM=2 =
6[(0,2)+(2,0)]+4(1,1)
28
(9.76)
and by orthogonality
1
DM=2 =
;
5
(9.77)
202
then
1
[(1,2)+(2,1)]+ 6[(0,1)+(1,0)]
14
[(2,2)+(2,2)]+4[(1,1)+(1,1)]+6(0,0)
70
2[(2,2)+(2,2)]+[(1,1)+(1,1)]2(0,0)
14
GM=1 =
GM=0 =
DM=0 =
(9.78)
and nally
1
S=
.
5
(9.79)
mn
p
mp
n
and then taking the trace, that is, setting m=n and summing over n, one gets
Tnn
= anx any Txy = xy Txy = Txx ;
in other terms, the trace Tmm is invariant. For a tensor whose cartesian
components Tmn = vm wn are the products of cartesian vector components
the trace becomes the scalar product Tmm = vm wm . Thus, in O+ (5) the
traces acquire particular meaning. The invariance remains obviously true with
any number of components; for instance taking the trace of
= amx any apz Txyz
Tmnp
203
one nds
= amx amy apz Txyz = xy apz Txyz = apz Txxz ,
Tmmp
thus Txxz is a vector. On the other hand a cartesian tensor of the form Txyz =
= amx any apz xy vz = amx anx apz vz =
xy vz is transformed as follows: Txyz
mn apz vz , and retains its form.
The and the T of Equation (9.74) are tensor components also for the
subgroup, except that since they are polar tensors the rule must be slightly
modied. In order to agree with the scalar product (9.8), the polar form of
the trace reads
()m Tm,m .
(9.81)
T rT =
m
(9.82)
Since the Tr operation commutes with the operations of the Group O+ (5),
many irreps of SU(5) are no longer irreducible. Vectors are sent to vectors and
are bases of an irreducible representation of O+ (5). Traceless tensors are sent
to traceless tensors and are bases of irreducible representations of O+ (5).
T
is traceless, while T
must be reducible. In fact, all the d2 states
m
mn
n
above belong to
of SU(5), but in O+ (5) the invariant 1 S is classied
in the (0, 0) irrep (the notation means: no boxes in either the rst and the
second line). The state with no electrons is already invariant, so this singlet
of SU(5), in O+ (5) one
is assigned the seniority number v = 0. From
can extract a traceless tensor which is classied in the irrep (2,0). All irreps
of O+ (5) are labelled by (1 , 2 ), 1 2 ; the parentheses denote the space
of traceless tensors. Since (2,0) cannot be made with fewer than 2 electrons
1
D and 1 F receive v = 2.
12
With
the 1-3 trace vanishes and the other two are equal in absolute
3
is a 5-component vector, so it deserves
value; clearly, vp = T r(1,2) T
mn
p
to be classied in (1, 0); it is obviously a doublet, so it corresponds to 21 D.
The other doublet states 23 P, 23 D, 23 F, 23 G, 23 H will correspond to the traceless
v
tensor T(2,0) T
mn5 p . This does not occur with less than 3 electrons
mn
p
nor with orbital angular momentum < 2. These states deserve the seniority
number v = 3 along with 43 P and 43 F that are classied in (1,1) and are also
unprecedented.
204
Problems
9.1. In a molecule of C3v symmetry, two electrons are in orbitals of the irrep
E. Find the space part of the singlet wave function of A1 symmetry (if any
exists) and the relevant CG coecients.
9.2. In a molecule of C3v symmetry, two electrons are in orbitals of the irrep
E. Find the space part of the singlet wave function of A2 symmetry and the
relevant Clebsh-Gordan coecients.
9.3. In a molecule of C3v symmetry, two electrons are in orbitals of the irrep
E. Find the space part of the singlet wave function of E symmetry and the
relevant Clebsh-Gordan coecients.
9.4. Build the character table for C4v
.
9.5. Consider the Cu++ ion, with one conguration 3d9 . How are the J =
3/2, 5/2 levels split by a square planar D4 environment?
Part III
(10.1)
(10.3)
Comparison with (10.2) is easier using the fact that Heisenberg operators
anticommute under T ordering , hence it holds that G2 (x1 , x2 , x3 , x4 ) =
T [ (x4 ) (x3 )(x2 )(x1 )] and
(z) (y) (y) (x) = ()G2 (x, y, y + , z)
(10.4)
208
i
H0 (x) ig (T ) (x, z) =
t
(10.5)
[i
H
dx lim+ lim
(x)
g (T ) (x, t, x , t )
i
0
t
t t r r
= i dxdyv(x, y) (x) (y)(y)(x) = 2iv .
(10.9)
The average of H0 can be obtained from Equation (4.26). So, we get the
exact ground state energy (lower sign for fermions):
i
E = H0 + V =
i + h0 (x) g (T ) (x, t, x , t )
dx lim+ lim
2
t
t t r r
(10.10)
H = H0 + H1 (t) =
k nk +
Vk,k (t)ak ak .
209
(10.11)
k,k
r
For igk,k
(t, t ) = (t t )[ak (t) , ak (t )]+ one nds
gk,k (t, t ) = (t t )k,k + (i)(t t )[ iak (t) , ak (t )]+ . (10.12)
t
t
Since
ia k = [ak , H] = k ak +
Vk,k (t)ak
(10.13)
k
r
gk,k
r
r
= (t t )k,k + k gk,k
Vk,p (t)gp,k
(t, t ) +
(t, t ).
t
p
(10.14)
(10.15)
r
No information about the lling enters the problem, and gk,k
is actually a
one-body quantity which does not depend on the lling. It is no harder to
calculate than a one-body wave function, even in time-dependent problems.
a
The same is true for advanced one, igk,k
(t, t ) = (t t)[ak (t) , ak (t )]+ .
For later use, we are interested in the time-ordered Greens function, dened
as usual by
(T )
(10.16)
igk,k (t, t ) = T ak (t)ak (t ) .
(T )
(10.17)
(10.18)
Let us calculate the t derivative of Eq.(4.19), taking into account that the
functions contribute
(t t )ak ak + ak ak = (t t )k,k .
Using (10.13), one nds
(T )
gk,k
t
(T )
p
(T )
(10.19)
210
This EOM is the same as (10.14), the dierence is in the initial conditions. I
stress that g (T ) depends on the Fermi level and can describe genuine manybody eects, so it is important to be able to express it in terms of the much
simpler g (r) . This is achieved as follows:
(T )
r
a
gk,q
(t, )gq,k
(10.20)
igk,k (t, t ) =
(, t )q (t, t ),
q
(10.21)
This solution was rst obtained by Ref. ([48]) by the Keldysh fomalism [84];
I found the simple derivation below and used the result to re-formulate the
theory of transport[63] (See Chapter 13.6.2). Since retarded and advanced
functions do not depend on lling, we may take the system empty, denoting
the vacuum by |0 , and write
r
ik (tt )
gk,k
,
(t, t ) = i(t t )0|ak (t)a (t )|0 = i(t t )kk e
k
a
ik (tt )
.
gk,k
(t, t ) = i(t t)0|ak (t)a (t )|0 = i(t t)kk e
k
Hence,
r
a
gk,q
(t, )gq,k
(, t) = k,k ,
(10.22)
r
a
fq gk,q
(t, )gq,k
(, t) = k,k fk .
(10.23)
Thus, the initial conditions are obeyed; the EOM are also satised because
r
(t, ): the (t ) term does not arise because < t,
they are obeyed by gk,q
but the required comes in from the derivative. Thus, (10.20) is readily
seen to satisfy EOM and initial conditions and is the exact solution.
10.2.1 Auger Induced Ionic Desorption: Knotek-Feibelman
Mechanism
Desorption (Section 6.2) is a process of emission of atoms, molecules or ionic
species that previously belonged to a surface or were adsorbed (i.e. chemically
bound to it). The amount of ions emitted depends on the electronic properties
of the species in a striking way. For example, if a Ag surface is bombarded
with Ar+ in the KeV range, some Ag + ions can be collected by a mass
spectrometer, but the ion yield is often much lower than the K + signal arising
from traces of K on the surface. This is because Ag + has a much lower
probability than K + of escaping without being neutralized.
When ionic surfaces are irradiated with X-rays one observes the desorption
of O+ or other positive ions that originally were anions[113]. When O
211
+
[Vk (t)a0, ak, + h.c.];
(10.24)
k,
at time t=0 two holes are created in the adatom orbital 0 and the consequent
desorption causes the time dependence. The band was assumed initially lled
and the amplitude for the two holes to be still on-site at time t is
N (t) = a0 (t)a0+ (t)a0+ (0)a0 (0) (t),
(10.25)
averaged over the hole vacuum. I solved the two-body problem by the EOM
method. Consider the correlation functions (averaged over the vacuum)
(t, ) =
k (t, ) =
(10.26)
(10.27)
(10.28)
(10.29)
(10.30)
one obtains
ia 0 = (t)a0 + k Vk (t)ak + U (t)a0 n0 ,
ia k = k ak + Vk (t)a0 (t),
(10.31)
i t
(t, ) = (t)(t, ) + k Vk (t)k (t, ),
(10.32)
212
we obtain
(10.34)
Moreover,
i
(t, ) = ( ) (t, ) +
Vk ( )k (t, )
+ iU ( )g(t, ) (, ),
(10.35)
(10.36)
+u( )
0
d u( )( ) (t, ).
(10.37)
+i
(10.38)
Finally,
2
(10.39)
H = H0 + H1 (t) + W =
k nk +
Vk,k (t)ak ak +
k,k
213
gk,k
(T )
(10.42)
(10.43)
appears. The interaction makes the problem hard and the equations do not
close any more. If one can be content with a mean-eld approximation, then
one can try
?
(10.44)
D,s,k (t, t ) T as, (t)ak , (t ) ns, (t)
?
which closes the equations again. Otherwise, one can generate an equation of
motion for D,s,k (t, t ) in order to truncate the hierarchy of Greens functions
at an higher level. When we cannot achieve the exact solution, one should
compare various approximations.
(10.45)
da+
= Ea+ + U a+ n ,
dt
da+ n
= (E + U )a+ n .
dt
214
This idea of the hierarchy can be generalized (see Ref.([49],[50][51]) and references therein). These authors call a composite operator, and write
a+
J(1)
(1)
a+ n (2)
J(2)
d
=
(10.46)
a = (3) , J = i dt = J(3)
J(4)
a n +
(4)
adopting a spinor notation (these are not spinors, they are just lists.) Anyhow, since the equation (10.46) is closed in this case, we may write
J(, t) =
(, ) (, t),
(10.47)
where for the sake of generality I imply a possible time dependence of the
operators. For H1site , the 44 matrix is block diagonal, since the rst two
entries are not connected to the others; each block is
E U
.
=
0 E+U
Following Ref.([49],[50][51]), we introduce
S, (t, t ) = (t t )0 |[ (t) , (t )]+ |0 ,
(10.48)
which is just i times the retarded Greens function (4.13), possibly averaged
over the Grand-Canonical Ensemble; we are using [, ]+ for Fermi particles,
but it would be [, ] for Bosons. The EOM read
i
S(t, t )
|t =t = [J(t), (t)]+ .
t
(10.49)
S(t, t )
|t =t ,
t
(10.50)
use Eq.(10.47) and the so called normalization matrix I = {I, } with elements
I, (t) = S, (t, t).
(10.51)
The nonzero diagonal blocks of I are
[a , a ]+
[a , a n ]+
1 n
=
for = +, .
[a n , a ]+ [a n , a n ]+ ,
n ns
Finally the EOM become
m = I.
(10.52)
215
(10.53)
(10.54)
the normalization matrix has site indices as well. From now on spinor indices
can be understood; we write the m matrix
m(p, q) = [J(p, t), (q, t)]+ .
(10.55)
Nop
(10.56)
with J is a nonlinear rest. In other terms, the set of operators is not closed,
and the spinors have more components than we can aord.
To keep the calculation manageable, one could trivially truncate the
spinors at Nop operators, throwing away the unwanted ones and J; the
matrix gives a solvable approximation. This is less accurate than the standard EOM method, which prescribes to replace the extra operators by some
approximation like (10.44).
However, there is a more clever alternative criterion for sorting out the
terms linear in in the r.h.s. of Equation (10.56). From the Nop -component
one can work out the truncated I matrix. The map (p), (q) Ip,q
sends two operators into a number like a scalar product in operator space. In
this Section I denote this scalar product by , and write
(p), (q) = I(p, q) = [ (p, t), (q, t)]+ .
Taking the scalar product of (10.56)
J(p), (s) =
Nop
(10.57)
Now in order to drop the last term we do not need to assume that J is
small; we make the milder assumption that it is orthogonal to the spinors.
Dropping J we nally obtain (10.52), that is,
216
= mI 1
(10.58)
and achieve a well dened, solvable approximation. In this way, the new
depends on those complicated operators that we are not considering explicitly. Although this is not a unique way to proceed, it is certainly appealing.
Fourier transforming to momentum space, the retarded Greens function is
approximated by
Nop
i (k)
,
(10.59)
g (r) (, k) =
Ei (k) + i
i
where Ei (k) are the eigenvalues of (k) and the spectral functions i (k) can
be derived[199] from I(k).
For a review on the applications to the Hubbard Model, see [198].
Problems
10.1. Verify (10.10) explicitly in the non-interacting limit.
10.2. Prove Equation 13.136.
10.3. Use the tight-binding method to calculate the retarded Greens functions of the tight-binding Hamiltonians in d=1,2,3 dimensions.
10.4. Dealing with X-Ray absorption and emission in Metals, Langreth[205]
considered the model Hamiltonian
H=
q aq aq + E0 b b +
Vqq aq aq bb
(10.60)
q
to mean that the product is over occupied spin-orbitals, and produces a Fermi
sphere in the thermodynamic limit. How can we nd the ground state energy
E0 of H? Standard perturbation theory fails unless V is small compared to
the unperturbed energy dierence; in practice, it fails almost always, since in
most interesting problems, the spectrum H0 is continuous.
Continuous spectra can be perturbed in such way that discrete spectra
arise. Suppose one wants to nd the bound states of the Hydrogen atom by
treating the Coulomb interaction as a perturbation: for any Z an innity of
bound levels exists, but there is no hope to get any sensible result with any
nite number of terms. The very existence of bound states requires poles of
1
= 1 + z + z 2 + . . . can only be
Greens functions to form. The pole of 1z
found at innite order. The formula (4.125) for E0 from the vacuum amplitude
(4.123)
(11.2)
R (t) = | UI (t) | = eiW0 t | eiHt |
is useful because we know R at least formally, recalling (2.36)
t
i
dt VI (t )
UI (t) = eiH0 t eiHt = T e h 0
t1
i t
i 2 t
=1+
dt1 VI (t1 ) + ( )
dt1
dt2 VI (t2 ) + ,
h 0
0
0
(11.3)
Note that in this section we use the telescope form (2.6) of the nested
integrals. To illustrate the method, I shall use the simplied interaction
V =
U (i, j, k, l)ci cj cl ck .
(11.4)
ijkl
218
In this way, we assume that only opposite-spin electrons interact, and this is
the simplication1 . Since |VI (t)| = |V | , the rst-order contribution
is
R
= i t | V | = i t
(1)
t
(t) = i
dt1 | V |
0
U (i, j, i, j)|ni nj |
(11.5)
ij
t1
j
l
Fig. 11.1. The rst-order contribution, with i=k, j=l. One says that i is contracted
with k and j with l. A more explicit denition follows shortly.
occupation numbers, and for each spin-orbital the creation operator is represented by a line leaving a vertex and the annihilation operator by a line
entering the vertex. Graphically, if we represent the up-spin operators as
the ends of the left line, and the interaction to the dashed line, this corresponds to the pattern of Fig. (15.4). The occupation number product over
the non-interacting | can be written as a vacuum average and using the
anticommutation rules we can separate the two spins: thus it factors:
|ni nj | = |ni | |nj | = ni nj
(11.6)
ijkl i j k l
t
t
U (i, j, k, l)U (i j k l ) dt1 0 1 dt2
0
| T [ci (t1 ) cj (t1 ) cl (t1 ) ck (t1 ) ci (t2 ) cj (t2 ) cl (t2 ) ck (t2 )] | .
(11.7)
More generally, we need to calculate the ground state average
M = | T [A1 (t1 )A2 (t2 )A3 (t3 ) . . .]|
(11.8)
where A1 , A2 , A3 . . . are interaction-picture operators - either creation or annihilation operators - dened on an orthonormal basis in the interaction picture and | is the Fermi sphere. This calculation can be worked out by
1
Later we can easily restore the full interaction by adding exchange terms.
219
writing down all the operators that make up the Fermi sphere according to
Equation (11.1); then this becomes an average over the vacuum |0 ; then,
by the anticommutation relations, we can bring the annihilation operators
to act on |0 to produce 0; all what remains from the anticommutators is
the result. However, we naturally ask if we can do something to reduce all
this groundwork. Fortunately, we can, in terms of contractions. A contraction of two operators A and B which are either creation or annihilation
interaction-picture operators, is dened as
A (t) B (t ) = | T [A (t) B (t )] | .
(11.9)
For this to be non zero one of the two operators must create and the other
annihilate the same one-body state, and the contractions yield (up to a sign)
the non-interacting propagators ( Equation (4.38))
igk (t) = T [ck, (t)ck, (0)] = eik t {(t)[1 nk ] (t)nk },
(11.10)
which propagates a hole for t < 0 and an electron for t > 0. For equal time
contractions one denes:
c (t) c (t) c (t)c(t) .
(11.11)
P
P () AP 1 AP 2
AP 3 AP 4 . . . ;
(11.12)
in words, M is the sum of the products ()P AP 1 AP 2 AP 3 AP 4 . . . of the
contractions; P is the permutation that takes from the initial expression
A1 A2 A3 . . . to the nal one, and ()P its signature. I propose an intuitive
proof2 .
As a warm-up, we calculate the ground state expectation value of the
product in the second-order term (11.7); for an easy start, we do so without
the time ordering T, and for the moment we let
M = | ci (t1 ) cj (t1 ) cl (t1 ) ck (t1 ) ci (t2 ) cj (t2 ) cl (t2 ) ck (t2 ) | .
Also, we assume provisionally that the canonical basis is the one of H0 eigenstates; then the time dependences are given by c-number phase factors as in
Equation (4.35), ck (t) = ck eik t . The operators come in pairs at each time
and since Fermi operators anticommute, one can lump all the spins on the
left without changing sign: we obtain
M = | ci (t1 ) ck (t1 ) ci (t2 ) ck (t2 ) cj (t1 ) cl (t1 )cj (t2 ) cl (t2 ) | .
2
220
(11.13)
The assumption that the canonical basis is the one of H0 eigenstates can be
removed, since anyhow H0 = H0 + H0 , and since H0 and H0 commute,
we can write c (t) = eiH0 t c eiH0 t and the time-dependent creation and
annihilation operators simply anticommute. While using Bloch waves to dene the Fermi sphere, we may wish to work with contractions in a site basis.
Next, we restore the T operation, and consider
M = | T [ci (t1 ) ck (t1 ) ci (t2 ) ck (t2 ) cj (t1 ) cl (t1 )cj (t2 ) cl (t2 )] |
= (t1 t2 )M12 + (t2 t1 )M21
(11.14)
M12 = | ci (t1 ) ck (t1 ) cj (t1 ) cl (t1 )ci (t2 ) ck (t2 ) cj (t2 ) cl (t2 ) | ,
M21 = | ci (t2 ) ck (t2 ) cj (t2 ) cl (t2 ) ci (t1 ) ck (t1 ) cj (t1 ) cl (t1 ) | .
Again, the operators come in pairs at each time, so in each term of the
sum ( 11.14) we can lump the spins on the left and factorize like above:
M12 = | ci (t1 ) ck (t1 ) ci (t2 ) ck (t2 ) |
|cj (t1 ) cl (t1 )cj (t2 ) cl (t2 ) |
M21 = | ci (t2 ) ck (t2 ) ci (t1 ) ck (t1 ) |
|cj (t2 ) cl (t2 ) cj (t1 ) cl (t1 ) |
and the whole time-ordered matrix element ( 11.14) breaks down into the
time ordered factors:
M = | T [ci (t1 ) ck (t1 ) ci (t2 ) ck (t2 )] |
| T [cj (t1 ) cl (t1 )cj (t2 ) cl (t2 )] | .
(11.15)
Thus, the time-ordering does not prevent the factorization, and the argument
works independently of the number of creation and annihilation operators.
Now we can breath a little bit, but can we do any better? We can actually
use the dierent spin-orbitals of any canonical basis as the dierent species,
but now we must keep track of the signs. Consider the matrix element for
spins. Particle i created at time t1 must be annihilated either by ck (t1 ) or by
221
ck (t2 ); so, two contractions contribute, but in each case the matrix element
breaks as before3 . Therefore the ground state average can be obtained as a
sum over permutations. However, there is more. In the above discussion, we
used no specic property of the ground state | ; we can average the matrix
element on any state in the same way (the contractions are then redened
accordingly). We can analytically continue Equation (11.12) from the real
axis to the verical track of Figure 2.2.2 b) (Sect.2.2.2). So, Wicks theorem
holds at nite temperatures as well; then stands for a thermal average.
Comments
This is a very general theorem that is often understated in books. It is obvious
that Wicks theorem also holds for averages over the vacuum |0 . By the way,
is itself a vacuum; for states above the Fermi level we can create electrons
like on |0 ; for occupied states the annihilation of an electron is the creation
of a hole. The transformation to hole operators c = b is canonical, or if you
like it is just a change in notation. So Wicks theorem works in both cases
for the same reason.
Wicks theorem for bosons works in the same way, and for the same reasons, except that the Bose operators commute under T and there is no sign
nuisance in this case.
11.1.2 Goldstone Diagrams
We represent the second-order terms diagrammatically by drawing two interaction lines, labeled t1 and t2 with t2 < t1 according to the scheme of
Fig. (11.1.2). These are time-ordered or Goldstone diagrams. Two diagrams
must be identied if they are topologically equivalent in the Goldstone sense,
that is, if they can be deformed into one another without changing the time
ordering. In this example with no interaction for parallel spins the vertices
on the left refer to spin up and there are no exchange terms. If we contract
only equal-time operators we get the a) diagram of Figure 11.1.2. All the nonpropagating lines (those that begin and end at the same time) must refer to
occupied orbitals (equal-time Wicks rule).
Diagrams consisting of two separated pieces, like a), are called unlinked.
Diagrams b),c) and d) are linked and b)and c) have the same value. The
contribution of diagram a) is obtained setting k = i, l = j, k = i , l = j in
R(2) (t) :
3
The possibility of contracting 4 or more operators with the same indices can
also be included. For averages on either |0 or , c c are either 0 or 1, so the Wick
factorization is granted since 0*0=0 and 1*1=1. So, these terms do not modify
the result. As emphasized in the Landau series[3], in the thermodynamic limit the
theorem holds for averages on any state, since taking the same index for two c c
factors means selecting a set of null measure.
222
j
i
t1
t2
k
l
time
R(a) (t) =
ij i j
+
| T [c+
(t
)
i+ 1 cj (t1 ) cj
t
0
dt1
t1
0
dt2
+
(t1 ) ci+ (t1 ) c+
i + (t2 ) cj (t2 ) cj (t2 ) ci + (t2 )] | .
(11.16)
2
t
R(a) (t) =
U (i, j, i, j)ni nj
U (i j i j )ni nj .
2
ij
(11.17)
ij
Here again we note that a line labeled i closing on itself, or tadpole, simply
contributes ni .
Diagram b) arises from the identications k = i, l = j, k = i , l = j and
to the contractions
+
+
+
c+
i+ (t1 ) ck+ (t1 ).ci + (t2 ) ck + (t2 ).cj (t1 ) cl (t2 ).cl (t1 ) cj (t2 ). (11.18)
and
(11.19)
c+
i (t2 ) ck + (t2 ) = (i , k ) ni
(11.20)
are immediate;let us consider the lines that represent propagators (see (4.37)).
The descending line yields
cj (t1 ) cl (t2 ) = (j, l ) eij (t1 t2 ) [nj (t1 t2 ) (1 nj ) (t2 t1 )]
= (j, l ) (i)gj (t2 t1 )
(11.21)
The - sign in front of (i)gj comes from the order of operators which is
opposite to the convention of ( 4.36 ).
223
(11.22)
In both cases, the time argument of g is the nal time minus the initial time
of the directed line. Since t2 < t1 , the contribution of the descending l = j
line represents a hole in a occupied state, while the for the ascending line is
an electron in an empty one. In both cases the directed line brings a factor
ig.
a)
i=k
b)
i=k
c)
d)
j=l
j=l
t1
j=l
j=l
i=k
j=l
i=k
j=l
t2
i=k
i=k
j=l
i=k
j=l
i=k
Thus, diagram b) ( like the identical one c)) yields the following expression:
R(b) =
ijkl i j k l
dt1
t1
0
dt2
(11.23)
ijli
0
dt1
t1
dt2
(11.24)
224
(11.25)
The term linear in t contributes to the ground state energy. Note however
that j is a hole, l is an electron,hence j l < 0 and the exponential in the
r.h.s. drops out as t (1 i). Problem 11.1 deals with diagram d).
Despite the heuristic appeal of cartoons showing the basic scattering
events between particles, the diagrammatic expansion does not look promising at all at this stage. The number of dierent diagrams and their complexity grows catastrophically with the order, and we need to go to order
to get useful results. Worse still, diagrams may defy intuition severely.
Their heuristic appeal can become very misleading, since the processes they
describe include counter-intuitive ones. Diagrams obtained from d) by appending tadpoles or bubbles freely to each line do belong to the expansion,
and must be included, even when the lines at a given time over-number the
electrons in the sample and/or violate the Pauli principle. No such principle
holds in diagrammatics, and several lines can bear the same quantum numbers. In one-particle problems, diagrams arise that appear to describe several
fermions propagating at a time.
11.1.3 Diagram Rules for the Thermodynamic Potential
Using imaginary time, the above results extend directly to the calculation of
the thermodynamic potential
where
= KB T lnZG
(11.26)
ZG = T re(HN )
(11.27)
is the grand partition function, the chemical potential and N the particle number operator. may be found as the sum of ring diagrams (see
=
Sect. 12.4.5) .The mean energy and particle number are related by E
each diagram: insert a factor rs|v|r s for each interaction (labeled by onen+1
body states, r and s entering and r,s leaving), a factor ()
2n n! and a (-) sign
1
for each loop 3) For each line labelled by r write a factor Gr (l ) = l
. This
r
225
factor is called
propagator. Next, sum over all internal labels and all frequen+
cies l with 1 l , including a convergence factor el 0 for loops. Examples
may be found in Ref.[55]. The T=0 version of the expansion
isobtained by
removing , replacing l by real frequencies and 1 l by d. Please
note that 1) simple diagrams may be easily computed without the diagram
rules directly from the T exp formula, 2) nobody ever computes very complex
diagrams, which are too many and too complicated to be of any use. The real
point is: how to avoid any heavy-duty use of the diagram rules.
= 2
Fig. 11.4. Combining diagrams. Those summed in the rst row consist of dierent
subdiagrams, while those of the second row are made up with like subdiagrams.
The only cheap and ecient way to produce and compute lots of highorder diagrams is by combining smaller ones. In a linked diagram one can
go from any vertex to any other one by following propagator or interaction
lines. The diagram a) of Figure 11.3 is an example of an unlinked diagram
made up of two simpler linked ones. Consider doing the same with any pair of
dierent connected diagrams. They give raise to dierent combined graphs for
R as shown in Figure 11.4. The diagrams on the l.h.s., top row, dier by the
order of interactions in time and are to be counted as dierent contributions
because the ordering is enforced by dierent functions; their sum is just the
product of the original diagrams. If however the same diagram is considered
twice, (bottom row and diagram 11.3 a) ) the contribution to R comes only
i } the set of linked
once so the result must be divided by 2. Denoting by {D
diagrams, consider a particular unlinked diagram of order n containing, say,
, D
, D
. Each of its n interaction lines labeled by times
3 linked parts D
tn < tn1 < < t1 belongs to one of the 3 sub-diagrams. Within each subdiagram, the ordering of the interaction lines is xed by the fact that it is a
226
here {ni } = n1 , n2 , is a list of non-negative integers specifying the num i , while the total number of parts is xed by the
ber ni of occurrences of D
Kroneker . This is just
1
ni (t)
D
(n1 , n2 , , nk , )!
(11.29)
i
p!
i
{np }
(
ni )!
where (n1 , n2 , , nk , )! = n1 !n2 !... is the multinomial coecient ; thus
Rp (t) =
Rp (t) ==
1
1
Di (t))p = (L(t))p
(
p! i
p!
(11.30)
Di (t) is the sum of all linked diagrams.
where L =
Thus we have obtained the simple but far reaching Linked Cluster Theorem:
R(t) = eL(t)
(11.31)
Every approximation to L(t) takes us to innite order in the series for R(t),
and a way to achieve real progress is open.
This exact result is the basis of the so called cumulant expansion4 . Now
our task is calculating L and the unlinked second-order diagram 11.3 a) must
be discarded, while the rst-order generates a partial series:
4
Note that a Linked Cluster Theorem exists[42] even if the interactions are
mediated by bosons; each sub-diagram must then be divided by its order. The
combinatorial argument is similar, but is modied by the fact that each line gets
two labels instead of one.
ln R(1) (t) = it
227
U (i, j, i, j)ni nj ,
ij
(iU )2
4
0
dt1
t1
dt2 e
2iV (t2 t1 )
U
= ( )2
V
1 + e2iV t + 2iV t
16
. (11.36)
228
t1
t2
t3
t4
Fig. 11.5. The fourth-order contributions to R(t). Note that two valence electron
lines propagate between t3 and t2 , although in this problem there is just 1 valence
electron. This illustrates that particle-number-violating diagrams must be included.
We get
t1
t2
t3
(iU )4 t
dt1
dt2
dt3
dt4 e2iV (t4 +t3 t1 t2 )
8
0
0
0
0
1 + 2e2iV t
U 4 5 e4iV t 4e2iV t
iV t
=( )
.
(11.37)
V
512
128
( )
Since the problem is simply solved exactly, we can check these results
by calculating R(t) and its expansion. One easily derives the Hamiltonian
on the basis of the U = 0 eigenstates and the ground-state-diagonal
matrix H
element its exponential:
=
H
U
2
V
U
2
U
2
U
2
+V
R(t) = e
iV t
2
{cos(
2iV
t
t
)+
sin( )}. (11.38)
2
11
229
Using (4.125) we compute the time derivative and let the exponential die out
as t (1 i); we obtain, in agreement with the exact solution,
U2
U4
U
+
+ O(U 6 )
2
8V
128V 3
W0 = V +
(11.40)
11.2.2 H2 Model
As a further example, we shall use the H2 model of Sect. 1.2.5.
igaa (t) =
1
{(t)eith t + (t)eith t }
2
(11.41)
6
1 5 ith t
e
(t) eith t (t) ,
2
6
1 5 ith t
e
(t) + eith t (t) .
2
The net second-order contribution comes from only 11.3 d). It is
iga,b = igb,a =
4 (iu)2
0
dt1
t1
16th
64t2h
(11.42)
where a factor 4 comes from the fact that each interaction line can be labeled a
or b indierently. The diagrams 11.3 b) and c) vanish because the terms with
the like interaction lines are canceled by those on interactions on dierent
atoms. In third-order there is no contribution because shifting any vertex
from an electron to a hole line changes sign to the diagram. For the same
reason, there is considerable cancelation in fourth order, the only surviving
terms being the one of Figure (11.2.2) and the one which results from the
exchange of t3 and t4 . Direct calculation shows that
it
itu
1 e4ith t
+ u2
2
16th
64t2h
e4ith t
itth
e4ith t
itth
.
1024
2048
512
C(t) = 2ith t
+
u4
t4h
+
5 e8ith t
8192
(11.43)
230
Fig. 11.6. One of the fourth-order diagrams that contribute to R(t) for the Hubbard H2 model. The interaction lines are labeled with time and site (si = a or b)
indices.
t
+
=
t
+
t
q
q aq aq + cc
gq (aq + aq )
(11.44)
has been used to discuss plasmon eects in core photoelectron spectra. The
removal of the core electron shifts the plasmon coordinates and the density
of states relevant to the spectrum is obtained from the correlation function
d it
e + |c (0)c(t)|+ ,
(11.45)
N () =
2
where + is the plasmon vacuum with the core electron present. N () is
obtained from the core propagator
g(t) = i+ |T c(t)c (0)|+ .
(11.46)
231
Langreth[39] noted that g (0) (t t ) is a simple phase factor for t < t while
g (tt ) 0 for t > t ; thus the ascending time line must be labeled t at the
bottom
and t at the top. From each diagram one can collect the same phase
factor: g (0) (ti tj ) = g (0) (t t ). Hence, g(t t ) = g (0) (t t )(t t ).
Erasing the deep hole propagator from the diagrams for g we nd those for ,
a succession of interaction lines ordered in a particular way; one can classify
these diagrams as linked and unlinked. Repeating the above argument one
nds that Linked Cluster Theorem applies,
(0)
(11.47)
where C is the only unlinked diagram, namely the second-order one: a boson
is emitted at time t2 > t and later re-adsorbed at t1 < t . Hence,
t
t1
2
(igq )
dt1
dt2 Dq (t1 t2 ).
(11.48)
C(t t ) =
t
(igq )2
t
dt1
q
t1
gq2
1 iq t eiq t
.
q2
(11.49)
a2
a3
exp[
an xn ] = 1 + a1 x + ( 1 + a2 )x2 + ( 1 + a1 a2 + a3 )x3 + . . .
2
3!
1
From a few diagrams for g that one can compute directly we can obtain
an equal number of an coecients that perform a particular innite resummation of the series.
232
a slightly lighter, discrete notation and the fact that for time-independent
problems the dependence is on a single time, namely, t = t1 t2 . In the
non-interacting problem with Hamiltonian H0 , the propagator is
kk
kk g0 (k,),
k + ik
(11.51)
(11.52)
ig(a, b, t) = 0 | T ca (t)cb (0) |0
averaged over the unknown interacting ground state, and the operators
are in the Heisenberg Picture. We use the standard denition (2.12) for the
Heisenberg operators but switch to the interaction picture, using (2.39)
AH = UI (t, t0 )AI (t)UI (t, t0 ),
in order to obtain g from an innite order series expansion of the evolution
operator in powers of V :
11.3.1 Adiabatic Switching and Perturbation Theory
To obtain g we do not really need 0 , as the denition seems to suggest. We
can resort to the trick of the adiabatic switching of V. Here we assume6 a
time-independent H and write
UI (t1 , t2 ) = UI (t1 , 0) UI (t2 , 0) = eiH0 t1 eiH(t1 t2 ) eiH0 t2
(11.53)
Assume that at time t = t0 in the remote past the system was in the unperturbed ground state , with energy eigenvalue W0 ; then at time 0 the
interaction picture state is a wave packet containing 0 :
|n n | eiEn t0 | (11.54)
UI (0, t0 ) | = eiHt0 eiH0 t0 | = eiWn t0
n
233
(11.55)
the one of the ground state dominates, and one is left with
UI (0, t0 ) | = |0 0 | eiE0 t0 | eiW0 t0 = |0 0 | UI (0, t0 ) | ; (11.56)
formally, (provided that the denominators do not vanish)
Im(t)
t0
Re(t)
t0
Fig. 11.8. The complex t plane with the tilted Gell-Mann and Low path.
|0 =
UI (0, t0 ) |
,
0 | UI (0, t0 ) |
0 | =
| UI (t0 , 0)
.
| UI (t0 , 0) |0
(11.57)
This is the Gell-Mann and Low Theorem [83]7 . Any expectation value
0 |A|0 on the interacting ground state can be obtained from non-interacting
ground state averages. One obtains:
| UI (t0 , 0) AUI (0, t0 ) |
;
| UI (t0 , 0) |0 0 | UI (0, t0 ) |
but along the tilted path, |0 0 | is equivalent to n |n n |; so
0 | A |0 =
(11.58)
The original proof reported by Ref. [2] is based on the perturbation series, so
it depends on its validity; also, it shows that numerators and denominators bear a
phase factor that diverges in the adiabatic limit.
8
Recall that with our convention the Schr
odinger, Heisenberg and interaction wave functions coincide at t = 0 and that at any other time the
change of representation is obtained by A (t) = I (0)| AH (t) |I (0) =
I (0)| UI (0, t) AI (t) UI (t, 0) |I (0) .
234
|UI (t0 , 0)UI (t, 0)ca (t)UI (t, 0)cb (0)UI (0, t0 )|
, t > 0. (11.60)
|UI (t0 , t0 )|
Using the Group property of U , the unitarity property UI (t, 0) = UI (0, t), and
the fact that under T the fermion operators anticommute, we shall manoeuvre
to obtain a single U in the numerator as well:
ig(a, b, t) =
|UI (t0 , 0)UI (0, t)ca (t)UI (t, 0)cb (0)UI (0, t0 )|
, t > 0.
|UI (t0 , t0 )|
(11.61)
In other terms,
ig(a, b, t) =
(11.62)
, setting S = UI (, ),
ig(a, b, t1 , t2 ) =
|S|
(11.63)
and expand in powers of V using the T exp formula (2.36). At each order
one obtains a sum of partial amplitudes, involving V and the bare propagator g 0 ; these expressions are best handled when represented as diagrams.
The key point is that dierent partial amplitudes give topologically inequivalent10 diagrams; at order n there is a nite set of possible topologies and all
correspond to partial amplitudes. In all diagrams, an oriented g 0 line enters
at a point b, goes through some interaction vertices and then reaches the exit
point a. Between two interaction vertices, the propagator line is labeled by
a spin-orbital: the one entering from outside will correspond to spin-orbital
b and the outgoing one to a. Dotted interaction lines start at each vertex;
at order n the diagram contains n interaction lines. A properly oriented and
labeled propagator line must pass by every vertex; each graph presents a path
which takes from b to a, and circuits attached to it (and possibly to other
circuits) by interaction lines.
The denominator of Equation (11.63) is like the vacuum amplitude in
(11.2) and yields all diagrams not connected to the main line; the only dierence is that in (11.2) the interactions are between times 0 and t and here
they can take place at any time. In the time representation (11.63), the
mathematical expression or amplitude of a disconnected diagram consisting of two parts is the product of the two amplitudes. Therefore, we may
9
235
(11.64)
a)
a,
b)
q,
b,
k,
q,
a
k,
q,
k
k,
k,
c)
q,
+
b
d)
Fig. 11.9. Diagrams for g: a) and b) exhaust the rst order, c) and d) are secondorder contributions.
the g 0 lines and the interaction lines by frequencies such that a and b have
frequency and at all vertices the sum of ingoing frequencies equals the sum
of outgoing ones. When momentum or crystal momentum are also good quantum numbers (translationally invariant or periodic systems) four-momentum
is conserved at each vertex. must pay attention to avoid double counting. To
see if two diagrams with an apparent correspondence between the vertices
are indeed topologically equivalent, one can start considering a path on one
diagram and the corresponding path on the other. If one meets the corresponding points in the same order in both diagrams, and this remains true
236
for all possible paths, the diagrams are topologically equivalent. For example,
the two diagrams in Figure 11.3.2 left are equivalent; I have added letters to
show that all vertices can be labeled in such a way that all paths correspond.
The orientation of arrows does not point to the evolution in time and each
propagator corresponds to electrons and holes, so the fact that one arrow
goes back in the second diagram is not meaningful.
Fig. 11.10. Left panel: two diagrams that look dierent but are the same. Right
panel: two ways to represent the interaction iVkpmn .
i
,
k +ik
(11.65)
for each electronic line with k = +0 for empty states and k = 0 for lled
ones (0 stands for a positive innitesimal, as usual). For every interaction line
labeled like in the left diagram a) include a factor
iVkpmn = i
d r
As we see in the right panel of Figure 11.3.2, what matters is the identity
of the ingoing and outgoing lines, and seemingly dierent pictures yield the
same expression. Multiply by (1) for each closed electron ring, sum over
intermediate
spin-orbitals and integrate over intermediate frequencies , , . . .
d
with d
2
2 . . .. This rule must be better specied when there are nonpropagating lines, like the tadpole. The contribution of the line (excluding
q,
(0)
nq
Fig. 11.11. A tadpole is a line that starts and ends at the same vertex.
d
ig0 (q,) = i
g0 (q, t = 0).
2
q
q
237
(11.67)
Since
(11.68)
(11.69)
where + is just after and just before. The transform must be taken for
t = 0 (the non-propagating lines simply have no time to propagate). Since
ig0 (q, t = 0 ) = n(0)
q ,
(0)
(0)
where nq is the unperturbed occupation number, the result is k nk .
(0)
Including a () for the circuit, the tadpole yields + q nq , which is
ready for inserting the q dependence of the interaction vertex. Let us see the
rst-order diagram (Figure (11.3.2), a)) and its value D[a)]: one obtains
0
0
0
D[a)] = ig (a,)ig (b,) (i)
(11.70)
Va k b k nk .
k
d
ig 0 (k, ) (i) Vk a b k .
D[b)] = ig (a,)ig (b,)
2
0
(11.71)
(11.3.2), c)) shows a second-order diagram with a couple of tadpoles inserted into a propagating line. In this case,
0
0
(11.72)
D[c)] = D[a)]ig (k,) (i)
Va k b k nk ,
k
where the new term in parentheses may be identied with the added tadpole
and the new ig 0 represents the new propagator piece.
So, we can calculate a subset of a diagram for later use, knowing that
it will appear in many other diagrams and bring some type of process into
play. As another example of the diagram rules, we calculate the pair bubble
238
d
ig0 (m,)ig0 (n, + ),
2
m,n
(11.73)
where the (-1) factor is due to the closed circuit and the 2 factor to the
sum over the spins of the circuit. We start evaluating the bubble with the
integral
d 0
g (m,)g0 (n, + ) =
2
1
d
1
2 m +im + n +in
(11.74)
by the residue method. When the poles are on the same side of the real axis
(m n > 0,) one closes the path on the other side and gets 0. Otherwise we
integrate in the upper half plane, and there are two cases: a) n occupied and
m empty b) m occupied and n empty. These represent respectively hole and
electron propagation. One nds:
ifn [1fm ]
n +m i0 case a
d
1
1
=
(11.75)
2 m +im + n +in
ifm [1fn ]
case
b
n +m +i0
So, we end up with
i0 () = 2
m,n
fm (1 fn )
fn (1 fm )
+i
.
+ m n i0
+ m n + i0
(11.76)
239
Fig. 11.12. Some of the lowest-order diagrams for g; there are two kinds in rstorder, but the variety grows with order in an impressive way.
the thick line, shows that as we proceed the number and complexity of the
diagrams grows in an inordinate way. However, although normally we cannot
obtain g exactly, we can use topology to make exact statements about g. All
terms except the rst have a factor ig (0) (a, )ig (0) (b, ), that is, an incoming
and an outgoing line. The stu in between is a self-energy part. The latter
does not have external lines, but is usually drawn with short ones to show
where they belong in the full diagram. Therefore the mathematical expression
for the self-energy part is just the one for ig (0) divided by ig (0) (a, )ig (0) (b, ).
Having computed a simple self-energy part, like the tadpole or the bubble,
one can conceive iterating it indenitely. The diagrammatic series is summed
240
energy part is called reducible if it can be split into two diagrams by cutting
a single propagator. The left diagram in Figure(11.4) is reducible, the right
one is not.
Fig. 11.14. Reducible (left) self-energy part and irreducible (right) self-energy
parts.
The Irreducible (or proper) Self-Energy is the sum of all the irreducible self-energy parts. The analytic expression of is i(), and the
function is also called self-energy. The rst-order self-energy part reads:
(11.77)
(i) (1) = (i)
Va k b k n0k
Vk a b k n0k .
k
The Dyson equation yields the exact diagrammatic expansion of the greens
function if the exact is known. Indeed, each diagram containing whatever
insertions in any number of times in any order comes out of by the
iteration of Figure 11.13 exactly once. Figure 11.13 reads:
ig = ig0 +ig0 (i)ig g = g0 +g0 g; g = g0 +gg0 ;
with x = (r, t) and dx = d3 rdt, these are
g(x, x ) = g0 (x, x ) +
g(x, x ) = g0 (x, x ) +
(11.78)
(11.79)
[i H0 (x)]g(x, x )
t
(11.81)
241
1
;
0 (k) (k,)
(11.82)
a
=
a
+p
a
q +
a
+
+
b
iab
,
a + ia
and the big dot between two lines , one ending with a label p and the second
starting with label q stands for for iVpq . In the Fano model one has only
the V0k matrix elements; to calculate the local Greens function g00 () one
can use two ways (see Figure 11.16).
(0)
(0)
() + g00
() 00 () g00 ()
g00 () = g00
242
i)
ii)
Fig. 11.16. Two ways to derive for the Fano model: i) the Dyson equation for
g00 ii) the Dyson equation for the g matrix.
One can write the Dyson equation using the self-energy i) for g00 (),
(0)
which is ready for inserting g00 lines; since
i() =
k
|V0k |
k + ik
(11.83)
g = g +g g
Vk0
k +ik g00
In the alternative method ii) we are seeking a matrix self-energy, where one
(0)
should insert not g00 lines, but any two lines.which is readily solved.
Another interesting example occurs in the theory of resistivity of metals;
the external potential is due to impurities. At second and higher order, the
self-energy is complex. One has the problem of calculating the Greens function and then averaging over a random impurity distribution; however, in the
dilute case, repeated scattering against the same impurity dominates. The
inverse quasi-particle lifetime 1 is given by Im; actually this can be taken
as 1 in Drude theory and is the dominant contribution to the resistivity at
low temperature, while at higher T phonon scattering becomes important.
243
g (1) including the eects of the direct rst-order Coulomb self-energy, i.e. the
tadpole (see Fig. 11.5a)). The self-energy is a matrix with elements (from
(0)
Equation 11.77) ab () = k Vakbk fk , that is,
+
b)
a)
ab () =
(0))
fk
d3 rd3 r
(11.84)
(0)
dr
(0)
fk |k(r )|2
;
|r r |
(11.85)
but this is nothing else than the electrostatic potential produced by the tadpole charge density. Had we considered non-interacting electrons in an effective external potential W (0) (r), we should have obtained the same ab .
Note that ab includes the eects of W (0) (r) exactly, as if we had found the
eigenstates {a(1) , b(1) , } of H0 + W (0) (r) and computed the new Greens
function g (1) , taking the matrix elements in the old basis {a, b, } We can
still improve the approximation using just the tadpole, but this time with
(1)
g (1) . The new correction ab can then be interpreted as if the electrons did
not interact, but moved in an additional potential W (1) (r). By iterating the
argument, one reaches the self-consistency, as shown in Fig. 11.5b) where
the exchange term has also been included, and the internal propagators are
dressed, renormalized, fully interacting propagators, shown as the thick lines.
This corresponds to the Hartree-Fock approximation. Hence, if one starts
with the Hartree-Fock basis as {a, b, } the series for starts with the
second order.
Second-Order
The proper (2) has two diagrams ( direct and an exchange).
The direct one is shown in Figure 11.18 a). It is convenient to introduce
the pair bubble (11.114), and to write:
244
b
b)
a)
i() =
d 0
ig (k,)(iVknbm )(iVamkn )(i0 ())
2
(11.86)
The calculation is an exercise in contour integration similar to the calculation of the bubble, and the contribution of Equation (11.76 ) to (11.86)
is
1
i
d
.
fn [1 fm ]Vknbm Vamkn
i
2 k + ik n + m i0
k,m,n
m,k,n
n +m k im
(11.87)
To get the exchange term from Figure 11.18 a) one cuts the lines raising from the bottom vertices and exchanges their upper ends, labeling in
such way that the exchange is done in the upper interaction resulting in
Vamkn Vamnk . The value of the diagram is obtained from (11.86) by
Vamkn Vamnk , (no closed circuit any more). Thus, the total second-order
self-energy is:
(2)
ab () =
[(1 fm ) fk fn + fm (1 fk ) (1 fn )]
m,k,n
(11.88)
245
Unlike the rst-order result, this correction is beyond Hartree-Fock, complex and -dependent. Starting from Hartree-Fock, we now obtain corrections to Koopman ionization potentials and electron anities and lifetimes
without resorting to the often prohibitive Conguration-Interaction computations. This is very useful in atomic and molecular calculations [4]. Typically
the corrections are of the order of 1 eV.
Fig. 11.19. By expanding the right-hand side, one nds an innite series of selfenergy corrections.
order of a series of more and more complex self-energy diagrams, and can
simplify enormously the task of summing the most relevant parts of the series
for .
Skeleton diagrams are those with no self-energy insertions. Since all the
internal propagators are dressed, to avoid double counting only skeleton diagrams are allowed. A few skeleton diagrams can replace an innity of selfenergy ones, if the self-consistency can be carried out by numerical iteration.
This can be a very good solution, depending on the choice of skeletons.
246
+
(11.89)
where is the scattering amplitude. The corresponding diagrammatic equation is shown in Figure 11.20. is the sum of all two-body diagrams such that
each ingoing line starts with an interaction, and each outgoing line leaves an
interaction. Internal lines can have all sorts of self-energy insertions.
1
3
G2
=
4
+
2
+
2
Fig. 11.20. The two-body Greens function G2 and the scattering amplitude .
The dressed incoming and outgoing lines belong to G2
Fig. 11.21. is the scattering amplitude; the heavy lines represent dressed propagators, the rst two diagrams shown are irreducible, the other two are not; J the
irreducible scattering amplitude.
In one can separate out the irreducible interaction diagrams, that cannot
be split in two by cutting only two dressed lines; let J denote their sum. All
terms in can be obtained by iteration from those of J, and this enables us
to write down a Dyson-like equation for .
Finally, the whole series is obtained from the Bethe-Salpeter equation
G2 (1234) = g(31)g(42)g(32)g(41)+ d5d6d7d8J(5678)ig(35)ig(46)G2(7812)
(11.90)
which is shown in Figure 11.23.
247
3
G2
1
+
7
G2
8
2
4
2
4
4
6
2
Fig. 11.23. The Bethe-Salpeter equation for the two-body Greens function.
2
=
1
1
Fig. 11.24. The relation of the self-energy to the scattering amplitude. Replacing
by its diagrammatic expansion one gets the expansion for . This diers from
Ref. [116], Chapter 10.6, since we assume that the Hartree-Fock approximation is
embodied in our bare propagator.
(11.91)
[i
H0 (1)]g(1, 1 ) d2(1, 2)g(1, 1) = h(1, 1 ).
(11.92)
t1
is a non-local potential that can have some local contribution (proportional
to (r r )), so there is some freedom in the denition of . One can decide
to include the Hartree potential VH (x) due to the charge ig(x, x+ ) as a local
part of or as a potential in H0 (x). We shall write the one-body term in
both the above equations as H0 (x) + VH (x), where the second term actually
comes from the local part of . Thus,
248
[i
(11.93)
We nd
h d2v(1, 2)G2 (1, 2|2+ , 1 ).
d2 [VH (1)(1, 2) + (1, 2)] g(2, 1 ) = i
(11.94)
We have obtained g, not just , since G2 comprises the incoming and
outgoing legs; a formal relation between self-energy and two-body function is
obtained by amputating the outgoing one, that is,
(1; 1 ) + VH (1)(1, 1 ) = i
h d2d3v(1, 2)G2 (1, 2|2+ , 3)g 1 (3; 1 ) (11.95)
where g 1 is the inverse of g in the matrix sense.
It is clear that the approximations for the two-body function and for the
self-energy cannot be chosen independently. The relation to the scattering
amplitude is given by the gure 11.23.
.
(x)
If we think of the integrals in discrete form, and the integrand does nor depend on derivatives of , the functional derivative is just a partial derivative
with respect to the value of in a particular space-time point. Functional
dierentiation is a powerful tool that we shall use in parallel with the diagrammatics. Functional derivatives with respect to the propagator g are
particularly easy because they undo the integrations which are prescribed by
the rules (see Figure 11.25. ) This remark will be useful later.
11
249
4
4
4
2
5
+
1
5
2
= (1, 2)
(1,2)
g(4,5)
(2)]|
|T [S(1)
,
|S|
(11.98)
since |S|
= lim,0 UI (, t3 + )UI (t3 + , t3 )UI (t3 , ), and
t +
UI (t3 + , t3 ) = 1 hi t33 H1 ( )d + one nds
S = lim UI (, t3 + )
,0
(x3 )
t3 +
i
dx
dt(x)(x x3 )(t t3 )
t3
i
UI (t3 , ) =
S(3),
h
(11.99)
|S|
|S|
(11.100)
250
g(1, 2)
= iG2 (1, 3|3+ , 2) + ig(1, 2)g(3, 3+).
(3)
(11.101)
This yields a new, useful link between g and G2 , but we are interested in
involving . We multiply by v(1, 3) and integrate over 3 (that is, over dx3 dt3 ,
of course). Since ig(3, 3+) = (3) is the density, we may introduce the
Hartree potential VH (1) = d3v(1, 3)n(3) and write
g(1, 2)
= i d3v(1, 3)G2 (1, 3|3+ , 2) VH (1)g(1, 2).
i
h d3v(1, 3)
(3)
(11.102)
Now replace 2 by 1 and 3 by 2. The result may be used again in (11.91)
H0 (1) g(1; 1 ) = (11 ) i d2v(1, 2)G2 (1; 2|2+ ; 1 )
i
t1
with added to H0 , and yields
g(1, 1)
; (11.103)
(2)
= (1) + d3v(1, 3)(3)
d2v(1, 2)
[i
H0 (1) VH (1)]g(1, 1 ) = (1, 1 ) + d2(1, 2)g(2, 1).
t1
again with VH changed to Vef f , yields
g(1, 2)
(1, 3)g(3, 2)d3 = i
h v(1, 3)
d3
(3)
(11.104)
(11.106)
251
now the rhs has a comfortable g factor on the right. Hence, post-multiplying
by g 1 we obtain the desired result
g 1(4, 6)
(1, 6) = i
h d(3, 4)v(1, 3)g(1, 4)
.
(11.107)
(3)
11.9.2 Polarization Bubble
The dielectric function
x ) is dened by the relation (using a notation
(x,
3
with x (r, t), dx d rdt,)
D(x, t) = dx (x, x , t t )E(x , t )
between displacement vector and electric eld which holds in linear media.
The external charges are the sources of D; hence, the eective potential
Vef f (r) due to an external source having bare potential is
(11.108)
Vef f (r) = dr dt 1 (r, t, r , t )(r , t ),
the inversion implied in the notation 1 is matrix inversion in the r, t, r
indices. Hence,adopting a lighter notation,
1 (1, 2) =
Vef f (1)
.
(2)
(11.109)
(11.110)
Fig. 11.26. The Dyson equation for the screened interaction W (heavy dashed
line) in terms of the Coulomb interaction (light dashed line)and the full polarization
propogator .
252
(11.112)
Here, is the sum of all the irreducible polarization parts that cannot be
split by cutting a single V line. Since the series for (1, 2) is symmetric
under a mirror reection that exchanges 1 and 2, (2, 1) = (1, 2), and also
W (2, 1) = W (2, 1); moreover the alternative form exists
W (1, 2) = v(1, 2) d(74)W (1, 7)(7, 4)v(4, 2).
(11.113)
The general structure of the diagrams for (Figure 11.27 ) shows that it can
be obtained from g and a vertex :
(1, 2) = i
h
g(2, 3)g(4, 2+) (341)d (34)
(11.114)
3
Fig. 11.27. the structure of , according to Equation (11.114), showing the points
where interaction lines can be inserted.
2
1
Fig. 11.28. The relation of the self-energy to the vertex function, after Equation
(11.115).
(1, 2) = i
h
253
(11.115)
in terms of the dressed interaction W and of the vertex . The fact that
the same function indeed appears in the expressions for and will be
apparent shortly.
11.9.3 The Vertex
We evaluate (11.111) and obtain a new form of Dysons equation (11.113)
for W , and a new formula to comparewith (11.114). Taking the functional
derivative of (11.110) Vef f (1) = (1) + d3v(1, 3)(3), one nds
Vef f (2)
g(4, 4+ )
= (2, 3) i
h d4
v(2, 4).
(3)
(3)
+
)
Since (11.114) says that we want two g factors, we evaluate g(4,4
by the
(3)
trick (11.106),
g 1 (5, 6)
g(4, 4+ )
= d(56)g(4, 5)
g(6, 4+ ),
(3)
(3)
and then substitute into (11.111): this gives the screened interaction
W (1, 2) = v(1, 2) + i
h v(1, 3)g(4, 5)
g 1 (5, 6)
g(6, 4+ )v(2, 4)d(3456).
(3)
(11.116)
The integral on the r.h.s. is a functional that begins with a bare interaction
v(1, 3) and ends with another bare v(2, 4). If we aim at (11.113) we must
convert the rst to a screened interaction: we need to screen it by and
we can if we screen as well. The correct way to do this is by changing
the independent variable from the external to the eective potential via the
functional chain rule
Vef f (7)
= d(7)
= d(7)1 (7, 3)
. (11.117)
(3)
(3) Vef f (7)
Vef f (7)
This produces
i
h
:
254
Here the square brackets contain W (1, 7) (see Equation 11.111); the result
agrees with Equation (11.113) if the under-braced quantity is (7, 4):
g 1 (5, 6)
g(6, 4+).
(7, 4) = d(56)g(4, 5)
(11.119)
Vef f (7)
This yields the following expression for the vertex of Equation (11.114):
(1, 2, 3) =
g 1 (1, 2)
.
Vef f (3)
(11.120)
(11.121)
(1, 2)
.
Vef f (3)
(11.122)
255
(1, 2)
.
Vef f (3)
(11.127)
d(45)
(1, 2)
g 1(6, 7)
g(4, 6)
g(7, 5).
g(4, 5)
Vef f (3)
1
(1,2)
Now inserting Equation (11.120) (1, 2, 3) = g
Vef f (3) we obtain the least
obvious of Hedins equations:
(123) = (1, 2)(1, 3) +
(1, 2)
g(46)g(75) (673)d(4567). (11.129)
g(4, 5)
Hedin[53] obtained these exact equations that formally determine selfenergy, polarization, vertex function and Greens function. Although they
were not solved exactly, they lie at the heart of powerful approximate methods for rst-principle calculations [54]. This equation lends itself to a diagrammatic interpretation (see Figure 11.10). Indeed, functional dierentiation may be understood as removing a propagator line from a self-energy
diagram, which then becomes a for-point function with 2 incoming and 2 outgoing lines. By the Hedin equation, all possible combinations and iterations
of such scattering diagrams give raise to the most general vertex.
256
4
5
6
5
7
1 2
1 2
Fig. 11.29. A skeleton Self-energy diagram contributing to (1, 2), the corresponding contribution to (1,2)
, obtained by deleting the g(4, 5) line (Figure 11.9), and
g(4,5)
the contribution to (2, 1, 3) arising from the Hedin equation (11.129).
Problems
11.1. Compute diagram d).
a)
b)
c)
k-q
m+q
+
k
Fig. 12.1. Diagram for g with a self-energy insertion involving 0 in Jellium.
d3 m
(2)3
fm (1 fm+q )
fm+q (1 fm )
h + m m+q + i
h + m m+q i
(12.2)
with
fm fm+q
,
Re0 (q, ) = 2 m h +
m m+q
Im0 (q, ) = 2 m fm (1 fm+q )(
h + m m+q ).
(12.3)
d3 q
(2)3
d 0
ig (k q, )(iVq )2 (i)0 (q, )
2
(12.4)
258
and since Vq = 4e
q2 , this diverges at small q. The electron gains self-energy by
exciting the medium and then re-adsorbing the excitations, but the process
runs out of control for the long-wavelength ones. What is going wrong at long
distances? It is the Coulomb interaction V , which is causing the divergence by
its long range, but should actually be replaced by a shorter ranged screened
interaction W .
Equation (11.112) is solved by Fourier transformation thanks to the translational invariance and becomes:
W (q, ) =
Vq
,
(q, )
(12.5)
(12.6)
in terms of the exact irreducible polarization part. The cheapest approximation prompts itself: it consists of using instead of the unknown the lowest polarization part; the resulting approximation is popular as the Random
Phase Approximation (RPA) based on
RP A (q, ) = 1 + Vq 0 (q, ).
(12.7)
As usual, simplicity brings extra benets; in this case, we can identify the
RPA as asymptotically exact in the case of high density (or perfect Fermi gas).
Dimensionally 0 = [E 1 ], that is, it is inverse energy, and any polarization
part is clearly
the same; to check this, recall that momentum integrals are
actually q summations and carry no dimension, interaction lines V and
frequency integrals d carry E and Greens functions g bring E 1 . Thus,
inserting a new V into a polarization
part to build
a more complicated one
brings a dimensionless factor q dVq g 2 q Vq g = [1]. However, if we
scale the density of the liquid, we are changing the Fermi wave vector kF ;
then for every interaction line Vq scaling like kF2 there is a q summation
and an energy denominator kF2 (each energy must be counted kF2 );
(kF3 )
dVq g 2 kF3 kF2 kF2 kF4 kF1 , and the factor scales like kF1 ,
thus,
q
that is, like the Wigner-Seits radius rs . Thus 0 is the dominant polarization
part at high density, and the RPA is good in that case. The second fraction
in Equation (12.2) may be transformed setting m + q = n, which gives
fm+q = fn , fm = fn+q ; now renaming with n m one nds
3
d m
0 (q, ) = 2
3 fm (1 fm+q )
(2)
1
1
.
(12.8)
h + m m+q + i
h + m+q m i
This was rst evaluated by Lindhard [52]. Writing q for
h
, a = q 2q , a+ = q + q2 ,
= 2E
F
q
kF
and setting:
mkF
[1
2 2 h2
1 + a
1 + a+
1
| [1 a2+ ] log |
|} ,
+ {[1 a2 ] log |
2q
1 a
1 a+
q2
q2
2
mkF
[1 a2 ], q < 2, q2 + q > > q2 + q
Im(0 ) =
2
2
4q
h
2,
q < 2, q2 + q > > 0
0
otherwise.
259
Re(0 ) =
(12.9)
A factor of 2 is due to the spin trace. For W = 0 one is left with a free-particle
problem and the zeroth approximation (0) such that its action on the plane
waves is (0) |k = f (Ek )|k . We consider the equation of motion
i
h
= [H, ]
t
(12.11)
for the one-electron density matrix in the presence of the perturbation W and
expand the density matrix = (0) +(1) + in powers of W . Linearizing the
equation of motion for the linear response (1) and taking matrix elements
between plane-wave states one readily arrives at
k|(1) |k + q =
t
(Ek Ek+q )k|(1) |k + q + [f (Ek+q ) f (Ek )]k|W |k + q . (12.12)
ih
Using k|W |k+q = W (q, )eit and dropping eit in all terms we obtain:
260
k|(1) |k + q = W (q, )
f (Ek+q ) f (Ek )
Ek+q Ek + h
(12.13)
2 f (Ek+q ) f (Ek )
.
Ek+q Ek + h
(12.14)
4e (ext)
n
(q, ) + n(q, )
2
q
(12.15)
where n(q, ) is the induced density. n(ext) (q, ) alone would produce a
potental Vq, and with a slight generalization of Equation (12.5) we write
4e (ext)
n
(q, ) = Vq, = (q, )W (q, );
q2
(12.16)
hence,
q2
((q, ) 1)W (q, )
(12.17)
4e2
and comparing with (12.12) we obtain the Lindhard dielectric function,
n(q, ) =
(q, ) = 1 +
(12.18)
2kF + q
2me2 kF
2me2 2
q2
) ln |
|.
+
(k
F
2 2
2 3
4
2k
h q
h q
F q
(12.19)
KT2 F
q2
(12.20)
4kF
is the Thomas-Fermi wave vector. Using (12.20), one nds
where KT F = a
B
that the screened potential of a point charge becomes a Yukawa potential
1r exp(KT F r). Actually, the asymptotic behavior of Fourier transform
reects the singularities [92]; at long distance the behavior of the screening
charge is oscillatory (Friedel oscillations):
n(r)
cos(2kF r)
.
r3
(12.21)
261
8e2
f (k ) [1 f (k+q )] (k+q k h) (12.22)
q2
k
k
qmax
Fig. 12.2. The longest and shortest q vectors for electron-hole excitations with a
given excitation energy. One starts with a Fermi wave vector k and must reach the
h.
energy surface at EF +
16e2
k+q k
f (k )
.
q2
(k+q k )2 (
h + i)2
(12.23)
ne2 E0
it
.
m 1i e
Thus,
262
J = E,
(12.24)
2
0
with 0 = nem . Assuming that everywith the conductivity () = 1i
thing depends only on z, Maxwells equations give:
2
4i
d2 E(z)
=
E 2 J(z).
dz 2
c2
c
Then (12.24) gives
d2 E(z)
2
4i()
= 2 1+
E(z).
dz 2
c
this is tantamount to say that the medium produces a refractive index nref
c
and c nref
with n2ref = = 1 + 4i()
. Thus we arrive at
=1
p2
( + i )
(12.25)
2
is the Plasma frequency. This is qualitatively correct for
where P = 4ne
m
simple metals, but looks very dierent from the Lindhard result.
Plasmons and the Lindhard dielectric function
Some resemblance of the Lindhard to the classical calculation is recovered by
a long-wavelength (q 0) expansion. The q 0 limit of Equation ( 12.23)is
obtained by inspection since the k q term in k+q k averages to 0 and
2
P
f (k ) = n2 , with the result that (q, ) 1 (+i)
2 . A more accurate
analysis readily gives:
(q, ) 1
hkF 2
P2
P2
3
(
)
q2 +
( + i)2
5 m
( + i)4
(12.26)
Plasmon modes are dened by (q, ) = 0, which is the condition for selfsustained oscillations (one can have W (q, ) nite with Vq, = 0. They are
collective modes of the electron liquid with
hkF 2 q 2
3
) 2 +
(12.27)
(q) = P 1 + (
10 m
P
With increasing q, eventually the plasmon branch enters the electron-hole
continuum and becomes unstable against converting into a pair. Actually they
are sharply dened as q 0 but are damped with increasing q; this Landau
damping is due to decay in multiple electron-hole pairs. Thus, the plasmons
dominate the high-frequency, long wavelength screening, while electron-hole
pairs in metals are slower and act as shorter distance (higher q) screening
263
3
(3 2 n) 3 2.21
kinetic t[n] = 10
rs2 Ry
exchange
0.916
Ry
rs
correlation 0.062 ln(rr ) 0.096Ry.
(12.29)
gQ
,
1 U gQ
(12.30)
where, as in (6.74),
g Q (z) =
q
1
.
z (Q q) (q)
(12.31)
264
of propagators; if the pair carries momentum Q and frequency , the propagators may be labeled Q q, and q, . According to the diagram
drules,
the interaction brings a factor iU and we must perform the q 2
sum
on internal labels. Doing the frequency integral by countour integration, we
can show that introducing an interaction and a couple of propagators brings
a factor U g Q (see Equation 6.69 with z = + i0).
+ ...
+ ...
Fig. 12.3. Ladder approximation for the scattering amplitude and for the selfenergy (the rst term is the Hartree contribution). Note that and are simply
related; there is a general link, as discussed in Section 12.5 below.
Then, the sum of the geometrical series gives the Kanamori result, that
we obtained in Sect. 6.3 by quite dierent means. Then the one-particle selfenergy at low density can be gained by Equation of Figure 12.3. This Low
Density Approximation (LDA) was rst proposed by V. Galitzkii [133].
H0 =
ij ci cj + c b b,
265
(12.33)
ij
with the rst term (in obvious notation) describes the band structure, b
creates the core electron;
H1 = U n0+ n0
(12.34)
introduces the local repulsion at site 0, and
H2 = Ucv (n0+ + n0 )b b
(12.35)
1;
(12.36)
(12.37)
+
i
+
i
0
0
(12.38)
T
0
dA))
(12.39)
266
The results were displayed numerically for the rectangular band model of
Section (6.2). It was necessary to remove the rst-order (tadpole) contribution
(1) = iU n :
(12.40)
it would make the site where the Auger decay takes place more repulsive than
the rest of the sites. Thus,
() = iU n +
d g0 ( )T ( + ),
(12.41)
iU
(2)
1 + iU g0 ()
(2)
, g0 ()
d
g0 ( )g0 ( ).
2
(12.42)
The integrals can be done with the help of the Lehmann representation.
When U is large enough to produce split-o states in the closed-band theory,
there is structure outside the continuum. One nds
Re() U 2 A( U )(T + + U )( + U ),
(12.43)
267
,
system,
f (t, t , p, k) =
,
c,c
(k)eiH(1)(tt ) HA (k) |
| HA
i[H(0)iop ]t
| ac e
ac | V (c , p) V (c, p)
(12.45)
Here, V (c, p) is the matrix element of the electro-magnetic Hamiltonian between the core-electron state c and the photoelectron state, H() is the
hamiltonian of the system with core electrons in the primary-hole state,
the operator op describes virtual Auger transitions, but is often replaced by
a constant
HA (k) =
M (k)a0 a0
is the operator that describes the valence hole creation at site 0 in the local
spin-orbitals denoted by greek letters; M (k) are Auger matrix elements.
Moreover, | is the true ground state with no core-hole ( = 1), and the sums
run over complete sets. For degenerate core levels a set of relaxed ground states
must be included. We considered a non-degenerate core state and assumed
that these sums are saturated by two main contributions, namely | = |
and | = | , where | is the relaxed ground state with the core-hole and its
screening cloud. In the case it is necessary one could easily extend the theory
including plasmon satellites and other excited states. One of the ingredients
is the core-hole Greens function gc (t) = i |ac ei[H(0)+iop ]t ac | . One
can observe the core density of states in Photoemission experiments as an
asymmetric line shape. We argue that the terms with
= are negligible,
since the factor
268
iH(0)(tt )
| HA
e
HA |
represents the evolution without the core-hole and it is unlikely that a screening cloud can be created taking from | to | if there in nothing to screen.
Let denote the inverse lifetime of the core hole and denote the inverse characteristic time of the electron-hole pairs screening the core hole
and contributing the asymmetry. We assume that both are small compared to the other relevant energies. thus, we may model the situation with
gc (t) = iei[ci( +)]t , where c is the core energy level, while
c (t) = i | ac ei[H(0)+iop ]t ac | ei tt
(12.46)
I = |M (k)|
d D(E )
2 ( c )2 + 2
(12.47)
ni ni
(12.48)
is the core hole-valence electron interaction at site i=0 (W > 0). Letting |
denote the ground state with no core-hole ( determinantal state in HF and
Local Density approximations) we looked for the ground state | with the
core-hole in the form
| + (a +
269
bq a0 aq ) |
1
(1 n)d lim
(h, E(q))|q .
h0
N q
(12.50)
(12.51)
The screening electron comes from the Fermi surface. Now within the twostep model one can obtain the Fermi golden rule expression for the relaxed
spectrum
1 2|M (k)|2
Im[
dtei(Ei0)t
gc(hhe) (, t)]
(12.52)
Irel (E) =
(1 n)d
0
where
gc(hhe) (, t) = (i)3 | T [a0 (t)a0 (t)a0 (t)a0 a0 (t)a0 ] |
(12.53)
(hh)
(e)
(he)
(h)
(, t) = gc (, t)gc (, t) + gc (, t)gc (, t)
gc
(he)
(h)
(h)
(h)
(e)
+gc (, t)gc (, t) 2gc (, t)gc (, t)gc (, t),
, , all dierent
(note the factor 2, which is needed to give the correct U=0 limit.) When two
indices coincide, we nd
(hhe)
(h)
(hh)
(h)
(, t) = (1 n2 )gc (, t) + gc (, t)gc (, t)
gc
(he)
(h)
(h)
(h)
(e)
+gc (, t)gc (, t) gc (, t)gc (, t)gc (, t),
1
The exact three-body wave function for a system of interacting identical particles can be obtained by a method by Faddeev [206] but here we were looking for
a simple approximation for the three-body Greens function.
270
(hhe)
(he)
(he)
(12.54)
(he)
d
(E
2i gc
)gch ()
(12.55)
For partially lled bands, a one-body contribution arises, along with the rest
of 3-body contributions and the 2-body ones in the unrelaxed contributions;
we called our approach the 1-2-3 theory. In was shown in Ref. [128] (see
Figure 12.5 ) that this theory qualitatively reproduces the experimental trend
for early transition metals.
12.3.3 Correlation in Early Transition Metals
Part of the intensity in Auger CVV and APECS spectra from transition
metals comes from the decay of core-holes that are unscreened in the initial
state; this is obtained by a two-body Greens function, as in the lled band
case. However, much of the intensity comes from core holes that are screened
when the Auger decay occurs.
Some of these spectral features are amenable to the one-body g, but the
rest requires[128] a two-hole-one-electron propagator, that is a 3-body function g (3) . In the strongly correlated case the approximation of the last Section
runs into trouble because the Herglotz property is not granted (the density of
states may turn negative.) The Core-Ladder approximation [127]is an extension of the ladder approximation to this problem that has several attractive
features. It is based on the idea of formally rewriting the problem in terms of
271
kinetic energy, eV
-6
Fig. 12.5. Thick solid line:schematic drawing of the experimental Ti L2 M45 M45 line
shape (see Ref. [160]) on a kinetic energy scale centered at the maximum. Dashed
line: the theory of Ref. [128] (dashed), with U =1.684 eV, band width w=7.14 eV
and n=0.254. Thin solid line: the closed-band theory result with the same U and w,
which predicts too low kinetic energy and is clearly not applicable in this case.
x
a)
b)
Fig. 12.6. a) typical second-order diagram for the 1 hole- 2-electron function g (3) (t);
b) the same diagram with a ctitious 3-body interaction that allows to carry on a
partial summation of the 3-line ladder series to innite order.
272
The third body is ltered through a x= i |i i|type projector, that is,
it is left undisturbed. In core problems, one can exactly dispose of the innite
summation in favour of a local projector. The same applies to localized splito states that occur at strong coupling. Thus one can base an approximation
that can be systematically improved by including more terms if needed. It
treats electrons and holes at equal footing, carries on a partial summation of
the perturbation series to innite order, and becomes exact both at weak and
at strong coupling (in the sense that it becomes equivalent to a full Ladder
approximation). A particularly interesting feature is that our approach grants
the Herglotz property, that is, the density of states is granted to be nonnegative. Indeed it is a common drawback of perturbation approaches that
potentially powerful diagram summations become untenable by the failure to
guarantee this zero-order requirement of positive probability. In this case we
achieve the result by proving that there exists a model Hamiltonian for which
our Core-Ladder series gives the exact answer. The theory naturally explains
the apparent negative-U behaviour of the early transition metal spectra.
g(x, x ) = g0 (x, x ) +
21
(12.57)
(12.58)
2
273
(12.59)
(12.60)
where
Z(1; 2) =
d
1 {(1;
1)g(
1; 2) g(1; 1)(1; 2)} .
(12.61)
(12.62)
1 2
g(1; 2)
2m
2=1+
1
(1)1 (1) 1 (1) (1) = j(1)
2mi
(12.63)
is the current density. Hence we end up with the continuity equation, which
is the number conservation law in a dierential form, provided that the right
hand side vanishes, and this requires
Z(1, 1+ ) = 0.
(12.64)
The exact Dyson equations (12.56,12.57) must imply the continuity equation
and (12.64). The best way to check this is to use the equations of motion
(10.7) involving the two-particle Greens function
2
d
+ 1 U (1) g(1; 1 ) = (1, 1 ) i d1V (1, 1)G2 (1; 1 |1+ ; 1 )
i
dt1
2m
(12.65)
and
21
d
U (1 ) g(1; 1 ) = (1, 1 ) i d1G2 (1; 1 |1+ ; 1 )V (1, 1 ).
+
i
dt1
2m
(12.66)
In this way, taking the dierence, we get
274
1
(1 + 2 ) (1 1 )g(1; 1 ) =
2m
= [U (1) U (1 )]g(1; 1 )
d
1[V (1,
1) V (1 , 1)]G2 (1; 1 |1+ ; 1 ). (12.67)
(n)
skel.
1
2
g
(n)
=
(12.68)
[g, ] =
n
n
2n
T r[g()n ()],
275
(12.69)
where the Tr operation now sums over spin and all one-electron labels and
integrates in d
. This leads to the following result.
Theorem 10. (Luttinger-Ward theorem) [55] The exact self-energy is given
by (note the order of arguments!)
(1, 2) =
.
g(2, 1)
(12.70)
A further theorem by Baym and Kadano [56] states that the continuity
equation and the momentum, angular momentum and energy conservation
laws are embodied in the functional;
Theorem 11. If (and only if ) a self-energy is -derivable, that is, comes according to (12.70) from some approximate , the approximation is conserving.
In other terms, even for approximate Equation (12.70) is equivalent to the
conservation laws.
The GW approximation is also -derivable and hence conserving.
1
2
+ 14
1
4
Fig. 12.8. Diagrams for and the corresponding . The rst-order example refers
to the Hartree-Fock approximation.
276
(12.71)
is invariant, since for any line entering a vertex there is another one leaving
it. In an innitesimal gauge transformation , g changes by the amount
g = [(1) (2)]g(1, 2). Then to rst order as a functional of g will
change by the amount
0 = = d1d2
g(1, 2) = d1d2 (2, 1)g(1, 2)
g(1, 2)
= i d1d2 (2, 1)[(1) (2)]g(1, 2). (12.72)
Exchanging the dummy variables in the second term, Eq.(12.72) yields
0 = d1d2 [(2, 1)g(1, 2) g(2, 1)(1, 2)](1)
(12.73)
The coecient of (1) must vanish identically. that is
d2 {(1, 2)g(2, 1) g(1, 2)(2, 1)} = 0.
(12.74)
(n)
2
g (0)
Fig. 12.9. Building a contribution (n) to the grand potential (occurring 2n times)
from a diagram for the self-energy. This looks very much like Fig. 12.7, but
please observe the dierences: propagators are bare, and the self-energy improper.
The grand potential is the sum of ring diagrams. One may work out the
grand-potential
and the ground-state energy
in parallel; just, in the former
case one uses 1 l , while in the latter case d. We use a short-hand notation valid for both cases 3 . Let (n) denote any such diagram of order n;
it may be thought of as the result of closing a self-energy diagram (improper,
in general) with a bare propagator. Mutatis mutandis, this is similar to the
3
Note however that some expressions are ambiguous if taken literally at T=0
and then one should take the T 0 limit at the end (see below).
277
procedure for building from the skeleton diagrams for the self-energy, since
there are 2n bare lines in (n) . J.M.Luttinger and J.C. Ward [55], derived
the diagram rules for the grand-potential in the frequency-momentum representation for Jellium. They wrote the contribution of order n, summed over
one-body states r, such that g (0) is diagonal,
(n)
1
1
(12.75)
T rg (0) ;
dgr(0) () r
n =
2n r
2n
here d stand for KB T times the summation over Matsubara frequencies
in the nite T case, and the Tr operation
implies summing over indices and
integrating over frequency. To compute n n one can use the fact that
an
a()d
n
n
an
a() =
=
.
(12.76)
n
0
n
n
Thus,writing r (, ) for at coupling constant ,
d (0)
1
= 0 +
gr () r (, ),
d
2 r
(12.77)
1
ln(1 + e(r ) ).
r
(12.78)
Using the solution to Problem 11.2 and the Dyson equation we end up with
d
1
gr ()r (, ),
= 0 +
d
(12.79)
2 r
0
which is a new expression for the ground-state energy in terms of the Greens
function, involving a coupling-constant integration. This implies
d
1
= T rg(, )(, ).
d
2
(12.80)
= 0,
g
(12.81)
278
5
6
1
T r ln( g01 ) + g =
++g
= ,
1
g
g
g0 g
(12.82)
(12.83)
Luttinger and Ward [55] further proved that, as the notation suggests, veries (12.80) and coincides with 0 for = 0; so, it is indeed the grand potential
(at nite temperatures) or the ground-state energy (at T=0). These results
have been generalized in the nineties; the Lund Group (C. -O.Almbladh and
which depends on g
U. von Barth) [136][137] have constructed a functional
and on the screened interaction W and is variational in both variables; they
by using simple approximations for
tested numerically the performance of
g and W and found that the results where competitive. The derivation in
Ref.[137] also points out the connection with Kohn-Sham theory (see below).
+ divJ = 0
t
(12.84)
notation, (x) = (x)(x), Ji (x) = i
(
. Let
2m
xi
xi ) (x )(x)
x =x
A1 , A2 denote Heisenberg operators such as fermion creation or annihilation
operators and an operator such as a density which commutes with A1 , A2
under Wicks T ordering. We know (see Problem 2.3) that
d
T {A1 (t1 )A2 (t2 )(t)} = T {A1 (t1 )A2 (t2 )(t)}
dt
+ (t t1 )T {[(t), A1 (t1 )] A2 (t2 )}
+ (t t2 )T {A1 (t1 )[(t), A2 (t2 )] }.
(12.85)
Following [140] (see also [1]) we set A1 (t1 ) = (x1 ), A2 (t2 ) = (x2 ), take
a (ground-state or thermal ) average, use the continuity equation to replace
and obtain
t T {(x1 ) (x2 ) (x)(x)} = T {(x1 ) (x2 )J(x)}
+(t t1 )T {[(x), (x1 )] (x2 )}
+(t t2 )T {(x1 )[(x), (x2 )] } .
(12.86)
h
( )G2 (x, x1 , x2 , x ) =
2mi
= i (4) (x x1 )g(x, x2 ) i (4) (x x2 )g(x, x1 )
279
t G2 (x, x1 , x2 , x) +
(12.87)
where (4) (x) = (3) (r)(t) is the 4-dimensional and the limit x x is understood. The continuity equation is an expression of charge conservation and
is associated by Noethers theorem to the gauge invariance. Ward identities
generally arise from invariance Groups of the theory.
280
Vext Vext
produces a new hamiltonian H , with a dierent ground state
wave function 0 and eigenvalue E0 ; by the variational principle
E0 = 0 |H|0 < 0 |H|0 = 0 |H + Vext Vext
|0 .
Thus,
E0 < E0 +
d3 x[Vext (x) Vext
(x)]n(x).
(12.89)
(12.90)
(12.91)
and summing the two one nds the contradictory result E0 + E0 < E0 + E0 .
Note that in the case of degeneracy the ground state density is not unique,
but this does not change the conclusion that the same n cannot be compatible
with two potentials. Thus, n Vext , or (in the DFT parlance) Vext (r) in an
unique functional of n(r). So, n Vext H and since in principle one can
solve the Schr
odinger equation, we may conclude that n 0 , E0 .
The functional E0 [n] is unknown: it would be the solution of all kinds of
ground state problems in a formula. We can think physically, and separate
out some obvious contributions. We set
1
2
dxdx
n(x)n(x )
|x x |
(12.93)
is the Hartree term; this shifts the problem to a new unknown universal
The electrostatic potential is
functional E.
n(x )
=
[E0 [n] E[n]].
(x) = Vext (x) + dx
(12.94)
|x x |
n
Since
E0
n
E
= .
n
(12.95)
281
Kohn and Sham [122] invented the fully quantum mechanical version,
just in time to take advantage of the computer revolution. To nd a quantum kinetic energy functional T [n], rst of all one must realize that it exists. As noted above n Vext H. Now, H determines 0 , E0 and also
T = 0 |T |0 . We cannot carry on this program because we cannot solve
the Schr
odinger equation for the interacting system. However for a noninteracting system S of electrons moving in some external potential VS we
can. Let TS [n] be the kinetic energy of a ctitious N-electron system S having
the same density as the real one; Kohn and Sham wrote
(12.96)
E0 [n] = TS [n] + dxVext (x)n(x) + VH [n] + Exc [n],
where Exc [n] is a new unknown exchange-correlation functional. Indeed, we
can apply to S the above argument showing that the density yields the external potential: n VS . Therefore, the required potential ensuring that S
has the exact density must exist; having separated out the kinetic energy of
S (that will dier from the exact kinetic energy of the interacting system)
we can write (12.96) and pretend we know all but Exc [n] potential that must
be approximated in some way, usually drawing from the Jellium theory at
the density n(x) prevailing at a given point. This is the Local Density Approximation. Among the popular approximations, I mention the interpolation
formula
1
4
4
3 3 3
n3
3
Exc [n] =
(12.97)
n 3 d3 r 0.056
1 d r
10
0.079 + n 3
and the extensions including gradient corrections
/
0
2
(n)|n(r)|
+
.
Exc [n] = d3 r xc (n)n(r) + (2)
xc
(12.98)
Exc
.
n
(12.99)
This contributes to the chemical potential, along with the electrostatic potential , since
E0
TS
==
+ + Vxc .
(12.100)
n
n
Thus in the ctitious system the electrons do not interact but feel a potential
VS = (x) +
Exc
,
n
(12.101)
which does not coincide with the screened potential but includes Vxc . One
must solve self-consistently the set of N equations, the Kohn-Sham equations
282
1
( 2 + VS )i (x) = i i (x),
2
(12.102)
N
|i (x)|2
(12.103)
i=1
from the determinant of the orbitals i (x). The exchange-correlation potential can be improved over the jellium estimate (12.29) by including the
generalized gradient corrections[138] and the accuracy of the results strongly
depends on the system. In metals, typical error may be 0.3 eV per atom
but for nite systems they are usually much worse. Bond lengths are usually
reproduced to better than 109 cm.
Further Developments
The density functional perturbation theory is a technique that allows to calculate e.g. phonon dispersion in solids[149]. I quote the following useful result
by Janak[146] without reporting the proof.
Theorem 13. In density functional theory,
E
= i ,
ni
(12.104)
that is, the derivative of the total energy with respect to the occupation number
of a Kohn-Sham orbital is equal to the eigenvalue of that orbital.
Vignale and Rasolt [119]formulated the current-density-functional theory,
an extension which is needed to include magnetic elds; the current couples
to the vector potential A and a term in A2 appears. A few pioneers [129],[130]
tried time-dependent versions of Kohn-Sham equations in atomic problems;
the results were rather encouraging, but a justication was lacking untill
Runge and Gross[131] found a suitable formalism. This was based on the
t2
283
(12.105)
(12.106)
Thus the Greens function gKS of the ctitious Kohn-Sham system obeys
1
1
gKS = gH + gH Vxc gKS gKS
= gH
Vxc .
(12.107)
(12.108)
(12.109)
Hence,
[gKS (xc Vxc )g]r =r,t =t+ = 0;
(12.110)
Problems
12.1. If as an approximate E[n] one keeps only t[n] (see Equation 12.28),
what kind of approximation is obtained?
12.2. Evaluate the diagram of Figure 11.2.2 with all si = a.
13 Non-Equilibrium Theory
286
13 Non-Equilibrium Theory
The operators in the Greens functions are in the Heisenberg picture; however
for any operator A we can switch representation by starting the evolution
from a golden age t0 when Heisenberg and Schrodinger pictures are the same,
according to
(13.2)
AH (t) = UI (t, t0 )AI (t)UI (t, t0 ).
There are four UI factors to expand in series in
g < (t, t ) = 0 (t0 )|H
(t )H (t)|0 (t0 ) =
0 |UI (t , t0 )I (t )UI (t , t0 )UI (t, t0 )I (t)UI (t , t0 )|0 ,
(13.3)
although we can reduce them to 3 since UI (t , t0 )UI (t, t0 ) = UI (t , t0 )UI (t0 , t)
and this is UI (t , t) by the group property. It would still be cumbersome to
287
expand the three UI factors, but we can do with just one expansion, since for
each operator A
AH (t) = UI (t, t0 )AI (t)UI (t, t0 ) =
t
t0
i
dt VI (t )
i
dt VI (t )
t0
t
= (T e
)AI (t)(T e
)=
t
dt VI (t ) AI (t)
= TC exp i
(13.4)
t0
where C is any oriented path in complex time through t0 and t , using the
generalized time-ordering TC along C that we met ins Section (2.2.1). Note
that AI (t) is under the action of TC that places it appropriately. In a similar
Im(t)
t0
Re(t)
Fig. 13.1. A contour on the complex t plane for inserting A(t) is a single
interaction-picture evolution.
way, we can read from left to right g < (t, t ) = 0 (t0 )|H
(t )H (t)|0 (t0 ) as
one story: the system starts at the golden age t0 , evolves to t, is acted
on by , then evolves to receive the action of at time t and eventually
it evolves back to the golden age. Physically, t can be before or after t. In
this story, we meet after because g < is dened with on the left of
. Thus along the path C = C1 U C2 , t precedes t , I write t <C t , and
Im(t)
t0
Re(t)
t
Fig. 13.2. The contour C for g < on
> the t plane with t <C t . Note that C can be
analyzed as a two-step path C = C1 C2 , C1 starts from t0 and returns there after
visiting t, and C2 starts from t0 and returns there after visiting t. The rst return
and restart from t0 can be avoided thanks to the group property.
288
13 Non-Equilibrium Theory
i
d VI ( )
i
d VI ( )
C1
C2
g < (t, t ) = TC1 e
I (t )
TC2 e
I (t)
i
d VI ( )
C
I (t )I (t)
= TC e
i
d VI ( )
C
= TC e
I (t)I (t ) , t <C t .
(13.5)
Because of the group property, the contour C is largely arbitrary. It can go
back to t0 between t and t any number of times, including 0. The terms
arising in the series development of the operators are ordered automatically
by TC with earlier times (on C) to the right. Moreover,
>
d VI ( )} I (t)I (t )
g (t, t ) = TC exp{i
C
= TC exp{i
d VI ( )} I (t )I (t) , t >C t
(13.6)
C
We can also use the same C in both cases, placing t and t in opposite orders.
As we know, the knowledge of both g > an g < gives access to the physically
important retarded and advanced Greens functions. We also need to dene
a time-ordered (on C) Greens function:
ig(t, t ) = TC H (t)H
(t ) = g > (t, t )C (t t ) g < (t, t )C (t t), (13.7)
(13.8)
following Langreth we must develop their combinations in series and in parallel which are needed to calculate diagrams in this theory. The combination
in series D = AB is dened by
d A(t, )B(, t )
D(t, t ) = AB(t, t ) =
C
(13.9)
289
(13.10)
d A(t, )ib< (, t )
<
C1
(13.11)
C2
t
t
6
d (i) a> (t, ) a< (t, ) b< (, t ).(13.12)
t0
Now we let t0 , and formally extend the integration to the full real
axis introducing a theta function. The result is
3
5
6
4
<
d [C1 ] =
d (i) a> (t, ) a< (t, ) (t ) b< (, t )
=
d ar (t, )b< (, t ) = ar b<
(13.13)
(13.14)
3
5
6
d ar (t, ) b> (, t ) + b< (, t ) +
6
4
5 >
a (, t ) + a< (, t ) ba (, t ) . (13.16)
(13.17)
However, the second integrand vanishes unless t > > t but then the in
front of the integral vanishes; therefore, we conclude that
290
13 Non-Equilibrium Theory
dr = ar br , da = aa ba .
(13.18)
(13.19)
(13.20)
(13.21)
(13.22)
(13.23)
Other forms are useful. Substituting a> (t t ) by (iar a> )(t t ), and
using b> + b< = ibr , we obtain
fr (t, t ) = i(t t )[ar b< a< br ];
(13.24)
>
dM = aM $ bM
d = ar b + a $ bM
d = aM $ b + a b a .
(13.27)
(13.28)
291
(13.29)
C
d VI ( )
(i)n
=
d1
dn V (1 ) V (n ).
n!
C
C
n=0
(13.30)
The diagrams are the same as in the Feynman scheme, but the time integrals
must be done on C, e.g. by the Langreth technique.
Example: Independent Particles
In independent-particle problems, V is a one-body operator. Lets write
V (t) = v(t) ,
(13.31)
is understood. In rst-order,
5
6
2
d v( )TC ( )( )(t) (t ) .
g = (i)
(13.32)
(13.34)
where + is just after and just before. Each contraction TC (t)(t ) =
ig 0 (t, t ) brings an unperturbed Greens function; each product of contractions
brings a factor ()P which is the signature of the permutation of operators
implied in doing the contraction. So, 13.32 becomes:
3
4
d v( ) ig 0 (, )ig 0 (t, t ) + ig 0 (t, )ig 0 (, t )
(13.35)
g = (i)2
C
292
13 Non-Equilibrium Theory
+
t
g = g 0 + g 0 g
(13.36)
(13.37)
Naturally, for t
= t one also denes
ir (t, t ) = [ > (t, t ) + < (t, t )] (t t ),
ia (t, t ) = [ > (t, t ) + < (t, t )] (t t).
(13.38)
(13.39)
ga = ga0 + ga0 a ga
>
g =g
4
0>
gr0 r g >
gr0 > ga
(13.40)
+g
0>
a ga
(13.41)
from now on, all operators are understood to be in the Interaction Picture, and
I drop the index I. Moreover Ill often understand
h to simplify the writing of the
equations.
293
(13.42)
However all the information is contained in g > and g < ; we need just
two equations. The equations for g (r) and g (a) are equivalent and provide
[ > (t, t ) + < (t, t )], so all we need is
f = g> g<
(13.43)
= > < .
(13.44)
(13.45)
We can also give a closed solution in terms of the self-energy. This equation
is a matrix equation in the indices t,t with the formal solution
"
5
61 ! 0
f = 1 gr0 r
f [1 + a ga ] + gr0 ga . .
(13.46)
1 gr0 r
61
moreover,
So,
5
6
= 1+gr0 r +gr0 r gr0 r + = 1+ gr0 + gr0 r gr0 + r = 1+gr r ;
5
1 gr0 r
61
f = [1 + gr r ] f 0 [1 + a ga ] + gr ga .
(13.47)
In summary, once is known, a clear cut procedure yields g. One can use a
and r to solve for ga and gr like in one-electron problems. The results and
are then used in Equation 13.47 for f , where the dependence on the Fermi
level enters. Eventually, from ga , gr and f one also obtains g < and g > .
Dynamical Independent-Particle Problems
The special case of independent electrons can be solved exactly but is far
from trivial due to many-body eects; this case also lends itself to include
interactions approximately in a dynamical-Hartree-Fock or time-dependent
density-functional scheme. In this case all the self-energy comes from the
connected rst-order diagram on the right of Figure 13.3, that is,
g =
d g 0 (t, ) v( ) g 0 (, t );
(13.48)
C
294
13 Non-Equilibrium Theory
g =
d
C
d g 0 (t, ) (, ) g 0 (, t )
(13.49)
(, ) = v( )C ( )
(13.50)
(13.52)
(13.53)
(t+ ) .
g c (t, t ) = iTC H (t+ )H
(13.54)
(13.55)
The traditional diagram expansion, that holds for constant V , is a special case
of the general one. The time sequence of Figure 13.5 in the interaction picture yields ()(i)U (, )U (, t )I (t )U (t , t)I (t)U (t, ) where a
295
(-) factor arises from the interchange of and . Including the other ordering of t and t, one obtains iU (, )T I (t)I (t )U (, ) with
the descending or backwards evolution operator U (, ) relegated on the
far left. The expectation value is taken at the golden age t0 over the noninteracting ground state (Chapter 1).
t
t0
Re t
Fig. 13.4. The Keldysh contour with t and t on the ascending branch.
t0
Re t
i
Fig. 13.5. The Keldysh contour with t and t on the ascending branch. The vertical
track is necessary for nite-temperature averages as in Figure 2.2.2.
gc (t, t ) = iTC H (t )H
(t ) = iTC H (t)H
(t ) ,
(13.57)
296
13 Non-Equilibrium Theory
i
g c (t, t ) = g > (t, t )(t t) g < (t, t )(t t )
(13.58)
g (t, t ) = iTC H (t )H (t+ ) = ig > (t, t ).
(13.59)
(13.60)
g c (t, t ) g + (t, t ) =
6
5
i g > (t t ) g < (t t) ig < [(t t ) (t t)]
6
5
= i g > + g < (t t ) = gr (t, t ),
(13.61)
Note that
(13.62)
On the Keldysh contour, C d = dt+ + dt ; so one can write all
integrals
on the real axis by introducing a matrix notation with the pattern
++ +
function
G(t, t ) =
gc g+
g gc
.
(13.63)
However, since we wish to rewrite dt as dt we must insert a
- sign for each integration in dt . This is conveniently done by inserting a
Pauli z matrix factor between any two matrix Greens functions in series
combinations. For example, the linked rst-order diagram due to a one-body
perturbation V (t) given by Equation (13.48) becomes
d g (t, ) v( ) g (, t )
g =
C
d v( ) G 0 (t, )z G 0 (, t ). (13.64)
Here
0
G (t, )z G (, t ) =
g c (t, ) g + (t, )
g (t, )
g c (t, )
g c (, t ) g + (, t )
g (, t ) gc (, t )
(13.65)
5
These are the correct equations, although in the literature gr and ga are often
confused.
297
where of course
=
c +
c
.
(13.68)
Again, this can be cast as in Sect. 13.3 in terms of advanced and retarded
functions. To this end, we use the unitary matrix
+
+
,
,
1 1
1
1
1 + iy
1
2
2
2
2
= 1 1
U=
, U =
(13.69)
12 12
2
2
2
leading to
+
U G U
gc g g+ +
gc gc g +g+
gc
2 +
2 +
c
c
c
g g +g +g +
gc
g +g g
2
2
0 ga
gr f
(13.70)
where f = if = g + + g and
U U
0 a
r
(13.71)
0 a
r
x =
r
a 0
,
(13.72)
298
13 Non-Equilibrium Theory
(13.74)
>
with Dmn
(t, t ) = m (t)n (t ) and so on, such that
>
<
iDmn (t, t ) = Dmn
(t, t )C (t t ) + Dmn
(t, t )C (t t).
(13.75)
(13.76)
Dmn
(t, t ) = in (t )m (t) = Dnm
(t , t).
(13.77)
(13.78)
a
Dmn
(t, t ) = i(t t)[m (t), n (t )] ,
(13.79)
299
with greek indices running over d and all the k states. Up to time t0
the system is in an eigenstate of H0 specied by occupation numbers n0 ;
this sets the initial condition of the problem. In some cases we wish to use
adiabatic switching to introduce Hh , as in desorption problems; in scattering
problems on the other hand the initial condition does not necessarily coincide
with the ground state oh H0 . The unperturbed Greens functions are readily
obtained.
0<
(t, t ) = c (t )c (t) = n0 ei (tt )
(13.81)
g
0>
g
(t, t ) = c (t)c (t ) = (1 n0 )ei (tt ) .
(13.82)
0
gr,
(t, t ) = i ei (tt ) ,
0
(t, t ) = +i ei (t t) ,
ga,
(13.83)
(13.84)
and we want
<
c
nd (t) = cd (t)cd (t) = gdd
(t, t) = igdd
(t, t + 0), t
(13.86)
300
13 Non-Equilibrium Theory
0
But gr,
and we can further simplify 6
0
0
gr, (t, t ) = gr, (t, t ) +
d gr, (t, )V ( )gr,
(, t ). (13.89)
The local (dd) component has a closed Dyson equation; indeed, using (in
matrix notation)
0
0
gr,dd = gr,dd
+ gr,dd
Vdk gr,kd
(13.90)
k
and inserting
one nds
0
gr,kd = gr,kk
Vkd gr,dd
(13.91)
0
0
+ gr,dd
r(d) gr,dd ,
gr,dd = gr,dd
(13.92)
0
Vdk (t)gr,kk
(t, t )Vkd (t ).
(13.93)
(13.94)
0
ga, (t, t ) = ga,
(t, t ) +
0
d ga, (t, )V ( )ga,
(, t ).
(13.88)
301
(13.95)
k
|Vdk |2
S (t) = i(t)
|Vdk |2 eik t ,
k + i
(13.96)
leading to
(13.97)
The dynamics of the virtual levels is factored from the dynamics due to
the time-dependent Hamiltonian. Using (13.83), the Dyson equation for the
retarded function reads
d1
gr,dd(t, t ) = ieit { eit (t t ) +
1
i1
S
d2 e u(1 )u(2 ) (1 2 )gr,dd (2 , t ;
(13.98)
(13.99)
t
P = gr,dd(, 0)|2
(13.101)
u(t) = et ;
(13.102)
1 A is the bond length. For any atom desorbing with a kinetic energy of some
eV one nds that h
302
13 Non-Equilibrium Theory
(13.103)
5
6
(13.104)
(13.105)
(13.106)
v0
1.
v 0, and any development in powers of fails despite the fact that
An evident weakness of the Lorentzian approximation is that P in (13.106)
is independent of the position of d with respect to the band continuum, while
the shape of the virtual level depends on it. The solutions of Equation (13.100)
(Ref. [47] ) show that P may be enhanced by orders of magnitude if the level
is near the band edges.
Including the Fermi Level
The general formalism is necessary when the band is partly occupied. f is the
quantity that depends explicitly on the occupation numbers, and the Dyson
equation for = 0 is solved by
f = [1 + gr r ] f 0 [1 + a ga ]
(13.107)
0
0>
0<
(t, t ) = g
(t, t ) g
(t, t ) = (1 2n0 )ei (tt ) .
f
(13.108)
The equation for f is a matrix equation in the time indices (t,t and intermediate times) and in those denoting spin-orbital states (d, k and intermediate states). Let me make such dependence explicit by the notation change
f [f ] (t, t ), and the like, when necessary; then, Equation (13.107) reads
8
The self-energy must have a factor (t) because it is retarded. However, the
product (t)(t) is ambiguous; here it is understood that (0) = 1 and therefore the
transform is S () = 2i.
f (t, t ) =
303
d1 [1 + gr r ] (t, 1 )
0
d2 f
(1 , 2 ) [1 + a ga ] (2 , t ).
(13.109)
(13.110)
0
d2 ga,
(t0 , 2 ) [1 + a ga ] (2 , t ).
(13.111)
This compact result completes the general solution of the many-electron problem with time-dependent Hamiltonian which is exact in the absence of interactions. The advanced and retarded are merely single-electron problems,
while the many body eects are accounted for by the equation for f.
Once we know gr , ga and f, since
2g < = g < + g > + g < g > = i(gr ga ) f,
(13.113)
<
(t). Applying Equation(13.113) to the vacuum,
we readily gain nd (t) = gdd
<
g 0 and i(gr ga) f = 0, hence (13.112) yields the rather unusual
expression of a matrix dierence as a matrix product
gr, (t, t0 )ga, (t0 , t ).
(13.114)
i(gr ga ) =
304
13 Non-Equilibrium Theory
<
2g
=
that is,
<
g
(t, t ) =
(13.115)
(13.116)
If an atom with an occupied d level scatters against the surface and the k
states of interest are empty, we shall consider n0d = 1, n0k = 0; then this is
essentially a one-particle problem and the result says nd (t) = |gr,dd (t, t0 )|2 . In
the case of F desorption from an alkali uoride, approximately n0d = 1, n0k = 0
in the hole picture. the many-body complications arise in the study of the
ionization probability of a positive ion hitting a metal surface. In this case,
n0d = 0 and
n0k |gr,dk (t, t0 |2 ,
(13.117)
nd (t) =
k
n0k
where
is the Fermi distribution. We can write the result in terms of the
local component gr,dd by using the appropriate component of the Dyson equation, namely,
0
gr,dk (t, t0 =
d gr,dd (t, )Vdk ( )gr,kk
(, t0 )
d gr,dd (t, )Vdk ( )eik ;
(13.118)
= ieik t
t0
305
<
(c ca ca c ) =
{g () ga
()},
(13.120)
J=
i
h
h 2 a
where a, stand for tho sites connected by the hopping T . The theory was
then extended to a continuous formulation. In both cases, the current-voltage
characteristics were obtained in terms of local densities of states calculated
on both sides of the partition. However, the left and right unperturbed subsystems cannot exist in insulation because the zero-order wave functions obey
306
13 Non-Equilibrium Theory
special boundary conditions at the idealized partition. The partitioned theory of transport through junctions was put on a more general foundation
by Feuchtwang[85], who obtained the current-voltage characteristics of onedimensional model systems, rigorous within the Schrodinger theory in the
non-interacting limit; he also used the Keldysh formalism but questioned
some technical aspects of the previous formulations.
The partitioned approaches manage to work out the theory without introducing the time by an artice. One considers that the left and right halves
of the system are already biased while an ideal partition prevents the electrical continuity. The adiabatic removal of the partition eventually allows a
stationary current to ow. Thus, the partitioned approach describes a situation that is quite dierent form the experimental one, when the connection is
there when the bias is applied. The static treatments were based on a crucial
assumption of equivalence with the physical situation; this assumption can
fail as we know today (see below).
Partition-free Dynamical Transport Theory
A time-dependent, partition-free framework was proposed in 1980 by the
writer[63]; the new scheme was found to be powerful to calculate the currentvoltage characteristics, but above all for the rst time allowed to obtain transient responses as well. The dynamical theory using the Keldysh formalism
and the equation of motion method is no harder than the static one and
yields much more information; it has not received much attention in the past
but it is gaining ground in recent years for its clear advantages. One crucial
point in favour of my approach is that it is partition-free: the whole system including the leads is in thermal equilibrium at some temperature until
the bias is applied. This is physically much more appealing, and reects the
experimental situation. I worked out the theory in parallel for discrete and
continuous models using time-ordered Greens functions and Equation (4.24)
for the current. In the discrete one-dimensional case, when the unperturbed
Hamiltonian is of the form
H0 =
cs (cs+1 + cs1 )
(13.121)
s=
Vs (t)cs cs
(13.122)
s=
5 T
6
eTs
T
lim gs,s+1
(t, t ) gs,s1
(t, t ); ,
h t t+0
(13.123)
307
this is physically equivalent if not identical to (13.120). where Ts is the hopping matrix element between site s and its neighbors. For the continuous
model,in obvious notation,
H0 =
k ck ck ,
(13.124)
k
H1 (t) =
(13.125)
k,k
and the current density was obtained by Equation (4.24). As we saw below
Equation (10.16), the equation of motion
i
T
T
T
gkk (t, t ) = kk (t t ) + k gkk
Vkp (t)gpk
(t, t ) +
(t, t ),
t
p
(13.126)
T
T
with the initial conditions gkk
(0, 0 ) = ikk (1 fk ), gkk (0, 0+ ) = ikk fk
T
T
( fk is the Fermi function) and gkk (t, t + 0+ ) gkk (t, t + 0 ) = ikk , are
satised by
(r)
(a)
T
gkk
fq gkq (t, )gqk (, t )
(t, t ) = i(t t)
i(t t )
q
(r)
(a)
(13.127)
provided that is earlier than t and t. This is essentially the Blandin et al.
solution[48] with a much simpler derivation. Here we need g T (t, t ) with t > 0,
r,a
r,a
and t just large than t, so we may set = 0. Writing gkk
(t) = gkk (t, 0), one
nds
e
h
(r)
(r)
Im
J (x, t) =
gx,q
(t) gx,q
(t).
(13.128)
m
q
The corresponding solution for the current owing at site {i, j, k} in a lead
oriented along the x axis in a 3d discrete model is
Jijk =
2eT
(r)
r
#
fq gijk,q (t)gi1jk,q
(t).
h
(13.129)
The solution is valid for any time dependence of V , but in order to make
the results more explicitly in Reference [63] I assumed H1 (t) = (t)V (x),,
with V (x) 0 deep in the left electrode and V (x) constant deep in the
right one. Now the retarded and advanced Greens functions can be computed
simply using a constant nal-state Hamiltonian.
In the non-interacting case, the current-voltage characteristics are obtained by taking the t limit by asymptotic methods[63] [92] and agree
with the results by Feuchwang[85]. It turns out that the current response is
308
13 Non-Equilibrium Theory
very indirectly related to the local Greens functions and is much more naturally written in terms of the asymptotic behavior of the wave function at long
distance in the free-electron wires. For small enough V , both the continuous
and the discrete models become ohmic and
J =
eV
.
(13.130)
It was shown in Ref. [63] and proved formally by Stefanucci and Ambladh [103] that for the non-interacting system the current-voltage characteristics obtained by the partitioned approach must coincide generally with
those of the partition-free approach; moreover, these Authors considered including correlation eects within the Time Dependent Density Functional
Theory (TDLDA) scheme. In strongly correlated systems and in the presence of bound states the proof does not apply and the current can oscillate
for long times, as in the Josephson eect . Examples are discussed in Ref.
[192] where Hubbard clusters produce Josephson supercurrents by the W=0
mechanism (Chapter 17). A new implementation of the time-dependent quantum transport theory within the TDLDA scheme has proposed quite recently
([102]).
Meir-WinGreen Weak-Coupling Formula
In 1992, Yigal Meir and Ned S. WinGreen [64] proposed a Laudauer formula
for the current through an interacting electron region. The system S is placed
between two non-interacting electrodes characterized by
<
>
() = 2i f () ( k ) , gk,k
() = 2i [1 f ()] ( k ) .
gk,k
(13.131)
They started from
e
d L <
L <
J=
VkL ,n gn,kL () Vk,n
gkL ,n () .
h
2
(13.132)
kL ,n
Here, gk<L ,n links the k state in the left electrode to the S state annihilated
T
T
<
gk,k
= Vk,m
() gn,m
() gk,k
() gn,m
() .
<
=
gk,n
<
gn,k
(13.133)
309
/
! T
"
! T
"0
<
T
<
T
;
gk,k
Vk,n Vk,m
gn,m + gn,m
gn,m
gk,k + gk,k
T
T
<
>
now we replace gn,m
+ gn,m
by gn,m
+ gn,m
, simplify according to
!
"
!
"
g< g> + g< g< g> + g< = g<g> g> g< ,
and use g > = g < + g (r) g (a) . Next, we perform the sum with k = d()
and introduce
m,n = 2
d()Va,n Va,m
,
(13.134)
a
where a adds some generality by summing over possible independent channels in the L electrode, like for instance angular momentum or spin.
J=
ie
h
d
/
0
5 r
6
(L)
a
<
m,n () f (L) () gn,m
gn,m
.
+ gn,m
(13.135)
m,n
The same formula with R in place of L will give the current entering from
the right electrode, which is the opposite in stationary conditions; thus one
can symmetrize the expression with
(L)
2
2
(L)
(R)
In the special case m,n (E) = m,n (E), that is when the couplings are
the same apart from a constant, the formula simplies to (see Problem 15.1)
J=
ie
h
(r)
(a)
m,n [gm,n
gm,n
].
(13.136)
m,n
where
=
Since
(L) (R)
.
+ (R)
(r)
(a)
m,n [gm,n
gm,n
] = 2iIm
m,n
(13.137)
(L)
(r)
m,n gm,n
,
m,n
this expression has a strong intuitive appeal if m,n can be taken real: contributions to the current arise from that slice of the density of states which
is occupied at the one side of the junction and empty at of the other
side. The density of states must be computed in the presence of the connections; however, the theory of [64] is particularly suited for small , when the
coupling to the leads is so weak that one can approximate g as the Greens
function of the isolated S.
Problems
13.1. Prove Equation (13.137).
Part IV
a0 b 1 0 0 . . .
b 1 a1 b 2 0 . . .
H=
0 b 2 a2 b 3 0 .
0 0 b 3 a3 b 4
... ... ... ... ...
(14.1)
(14.2)
1
|u0 ,
EH
E a0 b1
0
0
b1 E a1 b2
0
b2 E a2 b3
(E H)1 =
0
0
0
b3 E a3
...
...
...
...
(14.3)
1
...
...
0
.
b4
...
(14.4)
314
1
Now, zH
with z = E + i, = +0 is the Fourier transform of eiHt (t),
hence G0 (z) is causal; moreover it vanishes for z in all directions,
G0 (z) = G0 (z ) and is a Herglotz function, i.e. yields a non-negative density
of states n(E) = 1 ImG0 (z = E) 0. According to the standard rule, the
00 element is given by
D1 (E)
G0 (E) =
,
(14.5)
D0 (E)
E a0 b1
0
0
...
b1 E a1 b2
0
...
b2 E a2 b3
0
D0 = Det 0
,
0
0
b3 E a3 b4
...
...
...
... ...
0
0
...
E a1 b2
b2 E a2 b3
0
...
0
b3 E a3 b4
D1 = Det 0
(14.6)
.
0
0
b4 E a4 b5
...
...
...
... ...
Both are secular determinants of matrices representing chain Hamiltonians.
Similarly, we dene Dn as obtained by a crop of the rst n rows and columns;
it refers to a shortened chain with the rst n sites removed. Later, we shall
also consider shortening the chain the other way, dening N = D0 and
n , n < N the shortened chain with the rst n sites left. Now expand D0
by Laplaces rule along the rst line:
b1 b2
0
0 ...
0 E a2 b3
0 ...
.
D0 = (E a0 )D1 (b1 )Det
(14.7)
0
b3 E a3 b4 . . .
... ...
... ... ...
The b2 term in the rst line does not contribute, and
D0 (E) = (E a0 )D1 (E) b21 D2 (E)
which inserted into Eq.(14.5) yields
G0 (E) =
D1
D1
1
=
=
.
2
D0
(E a0 )D1 b21 D2
(E a0 ) b21 D
D1
Then,
D2
1
=
3
D1
(E a1 ) b22 D
D2
and iterating,
G0 (E) =
1
E a0
b21
Ea1
315
(14.8)
b2
2
b2
3
Ea2
Ea3 ...
1
.
E b2 G0 (E)
(14.9)
(4b2 E 2 ) 4b2 E 2
n(E) =
.
2b2
14.1.2 Greens Function of any System
Consider an arbitrary Hamiltonian H, possibly representing a complicated
many-body system, and any normalized state u0 ; if this is an eigenstate we
are nished, otherwise we may always write :
H|u0 = a0 |u0 + b1 |u1 , u0 |u1 = 0, u1 |u1 = 1.
(14.10)
1
(H a0 )|u0
b1
(14.11)
(14.12)
where v denotes the norm of v. This can be computed from the given H
and u0 ; now we also know u1 and we can proceed with
H|u1 = a1 |u1 + b1 |u0 + b2 |u2 , u0 |u2 = 0, u1 |u2 = 0.
(14.13)
Here, b1 is the same as above since H is Hermitean and u2 is taken to be normalized. Thus, a1 = u1 |H|u1 , and b2 |u2 = (H a1 )|u1 b1 |u0 . Imposing
that u2 is indeed normalized, we obtain, since the phase of b2 can be chosen
at will,
(14.14)
b2 = (H a1 )|u1 b1 |u0 ;
again, we can compute that, write
u2 =
and expect the next step
1
[(H a1 )|u1 b1 |u0 ]
b2
(14.15)
316
(14.16)
with
an = un |H|un
(14.17)
(14.18)
(14.19)
317
which is a bar diagram but converges to the exact result in norm. One can
smooth it by tting with a continuous function and then dierentiate to nd
an approximate n(E); the results are much better than before, and at least for
one-electron problems converge quickly almost everywhere. Interestingly, for
regular solids there are always special intervals centered at E values where the
approximation fails badly; these singular values, where n is a bad function, are
dn
is discontinuous.
band-edge singularities or Van Hove singularities , where dE
If one is interested in the overall line shape of some spectroscopy, this may
be of minor importance, since such singularities tend to be masked somewhat
by the experimental broadenings. However there is a message there: the longtime behavior of the Fourier transform of any function of frequency depends
[92] precisely on such details, and the transform is a correlation function,
which can be observed in appropriate experiments. We know from Section
0.2
0.2
0.2
Fig. 14.1. The rectangular n(E) with W = 0.5 and a broadening = 0.05 (dashed)
compared with the continued fraction approximations using N=5 denominators
(left), N=10 denominators (centre), N=40 (right). The continued fraction coecient
are discussed in the next Sections.
318
0 (E) =
1
E 0
12
E1
E2
(14.21)
2
2
2
3
E3 .........
2
N
1
2 t (E) .
E N 1 N
N
Here, tN (E) denes the tail of the continued fraction. The idea is: pick the
tail and attach it to G. Doing this by brute force is not rewarding, while
deepening the mathematics is a far better idea.
Pade Representation
Orthogonal Polynomials
It is evident that the continued fraction truncated at level N is the ratio of
two polynomials. To nd them, consider the linear chain eigenstates,
N
fn() |n
(14.22)
fn fn = , ,
(14.23)
| =
n=0
that satisfy
N
n=0
n=0
= E ()
fn() |n , b0 = 0.
(14.24)
n=0
N
N
fn bn+1 |n + 1 =
n=0 fn1 bn |n , and
n=0 fn bn |n 1 =
f
b
|n ,
(f
=
0),
one
nds
n+1
n+1
N
+1
n=0
Since
N
n=0
()
()
(an E () )fn() + bn+1 fn+1 + bn fn1 |n = 0.
(14.25)
n=0
()
()
f1 = 0.
(14.26)
319
f0 0| does not vanish (otherwise fn 0 ) and we may dene N independent functions
f
Pn (E () ) = n ,
(14.27)
f0
having the same r.r.
(an E () )Pn + bn+1 Pn+1 + bn Pn1 = 0,
P1=0 .
(14.28)
but P0 (E) = 1; all the others are determined by the r.r. and depend on the
a, b coecients of the previous sites. Evidently, they are just polynomials in
E. Changing the sum
N
fm
fn = m,n
(14.29)
to an energy integral,
m,n =
N
dE(E E () )fm
fn
(14.30)
dE
N
(14.31)
m,n =
dE n(E)Pm
(E)Pn (E)
(14.32)
with a weight n(E), the local density of states at |0 (compare with Section
5.1.4).
Polynomials, Secular Determinants and G
The secular determinant N (E) consists of the rst N rows and columns of
D0 (Equation 14.6)
0
0
...
E a0 b1
b1 E a1 b2
0
...
0
b2 E a2 b3
N = Det 0
;
(14.33)
0
0
b3 E a3 b4
...
...
...
. . . . . . N N
let us consider then the succession of determinants n (E) and expand
n+1 (E) using the last row:
320
n+1
... ...
...
...
...
. . . E a3 bn2
0
0
0
= Det . . . bn2 E an2 bn1
...
0
bn1 E an1 bn
...
0
0
bn
E an n+1n+1
... ...
...
...
. . . E a3 bn2
0
= (E an )n + bn Det
;(14.34)
. . . bn2 E an2 0
...
0
bn1 bn nn
n
bn .
(14.35)
(14.36)
i=1
Thus, we have got D0 = N (E) for the N -site chain in terms of the Pn and
in view of (14.5) a new expression for G0 is within reach if we can also express
D1 (E). However, D1 (E) is the counterpart of D0 (E) in a chain where site 0
has been cut o. The polynomials of that chain satisfy the same r.r.(14.28)
but they vanish on site 0. We can specify them as follows:
Q0 = 0, Q1 (E) = 1, (an E)Qn + bn+1 Qn+1 + bn Qn1 = 0.
(14.37)
Pn and Qn are the independent solutions for the N site chain that correspond
to dierent initial conditions. Therefore,
D1 (E) = QN (E)
N
bn .
(14.38)
i=2
G0 (E) =
QN (E)
.
b1 PN (E)
(14.39)
321
pick up tN (E) from the terminator continued fraction, but we are not so
silly to use it directly. For the truncated terminator, we write, in place of
(14.39,14.28,14.37)
N (E)
(N )
0 (E) =
.
(14.40)
1 N (E)
where s are the counterparts of the Qs in the terminators chain; the
polynomials,
(n E () )n + n+1 n+1 + n n1 = 0,
1=0 .
(14.41)
(14.42)
(14.43)
1 E N 1 N 1 (E) N 1 N 2 (E)
.
1 E N 1 N 1 (E) N 1 N 2 (E)
(14.44)
0 (E) =
2
tN (E)]N 1 (E) N 1 N 2 (E)
1 E [N 1 + N
.
2 t (E)]
1 E [N 1 + N
N
N 1 (E) N 1 N 2 (E)
(14.45)
1 N N tN (E)N 1
.
1 N N tN (E)N 1
(14.46)
N 1 0 N
1
.
N 1 N 1 0 N 1
(14.47)
1 QN bN tN (E)QN 1
.
b1 PN bN tN (E)PN 1
(14.48)
This method can give excellent results already with N a few tens if the
terminator has the correct edges of the continua with the right edge and Van
Hove singularities.
322
14.1.4 Moments
The above method is in a way a rened version of the moments method [60],
which is itself worth considering for its conceptual and practical importance.
Let |0 be any state of the system, and
1
n() = 0|( H)|0 = ImG()
(14.49)
the local density of states. Even for complicated H we can obtain useful
results by computing the moments
n
n = 0|H |0 =
d n n();
(14.50)
if we could obtain them all and sum an exponential series we would build
the Fourier transform of (14.49). Now, 0 = 1, 1 is the center-of-mass of
the virtual level, 2 informs us about its width, 3 characterizes its skewness, and with increasing n ner details are revealed. The moments allow
computing
the Greens function: if Re z is larger than the eigenvalues of H,
k
G(z) =
k=0 z k+1 . The short time behavior of n(t) depends on the rst
moments, that we can compute; the long-time n(t)is often needed and even
the asymptotic trend is of interest, but that information must be sought elsewhere. Letting N denote the number of moments that we can compute, we
face the problem of using them in the best way. One possibility would be
choosing a functional form which depends on N parameters and imposing the
values of N moments. For instance, if the edges of the continuum are known
one can pick a polynomial or an expansion in Tchebychev polynomials. However, the Van Hove singularities will prevent the uniform convergence of the
procedure. A more powerful method exists. By the M + 1-site chain Hamilp
)00 by
tionian1 HM of the form (14.2) one computes the moments p = (HM
matrix multiplication; from p , p = 1, N one can deduce the ak and bk
coecients for k N 1. In fact, 1 = a0 ; inserting into 2 = a20 + b21 we
have one unknown and solve immediately obtaining b21 = 2 21 . The third
moment is
3 = a30 + (2a0 + a1 )b21 ;
inserting the previous results we obtain
a1 =
31 + 21 2 3
.
21 2
323
32 + 21 2 3 23 21 4 + 2 4
.
(21 2 )2
General explicit formulas are also available from the theory of Pade approximants [157].
In this way we solve the inverse of Haydocks problem: given a normalized
Herglotz function n(), to determine a semi-innite chain Hamiltonian such
that n() is its density of states. For a symmetric n(), like those arising from
bipartite lattices (Section 4.3.1 ) we may take ak 0. For the rectangular
band model (14.20) one readily determines the coecients (Problem 14.2);
in this way I performed the comparison between the truncated fraction and
the exact density of states following Equation(14.20).
L | |
(14.51)
which shows how the and congurations are entangled. The particles of
one spin are treated as the bath for those of the opposite spin: this form also
enters the proof of a famous theorem by Lieb[37] discussed in Section 18.6.
In Equation (14.51) L is a m m rectangular matrix.
Schr
odinger Equation
Let K denote the kinetic energy m m square matrix in the basis {| },
()
()
and ns the spin- occupation number matrix at site s in the same basis (ns
2
If it is sparse, one can store and manipulate the nonzero elements Hij together
with i and j; this procedure is annoying and leads to a slowing down of the computations. Moreover, in many problems of interest H is not sparse.
324
s,s
(K ) and the like, one nds that
H0 | =
L [| | (K ) + | | (K ) ]
(K L) | | +
5
(LKT ) | |
6
T
K L + LK
| |
(14.52)
U (s, s )
s,s
| |ns | ,
L, | | (ns ) (ns ) ;
(14.53)
, ,
U (s, s )
s,s
(ns Lns ), | |
(ns Lns ), | | .
(14.54)
(14.55)
s,s ,
s,s
For illustration, below I adopt the on-site Hubbard interaction, when this
reduces to Liebs rule [37]
()
n()
(14.56)
H[L] = [K L + LKT ] + U
s Lns .
s
In the absence of magnetic elds, KT can be taken real and symmetric, hence
we drop the transposition sign. In particular in the Sz = 0 sector (M = M )
()
()
and K = K , ns = ns . The action of H is obtained in a spin-disentangled
way, in terms of operators acting in the spin subspaces. The generality of
the method is not spoiled by the fact that it is fastest in the Sz = 0 sector,
because it is useful provided that the spins are not totally lined up; on the
other hand, Sz = 0 can always be assumed, as long as the hamiltonian is
SU (2) invariant. For example, consider the Hubbard Model with 2 sites a
and b and 2 electrons (H2 molecule, Section 1.2.5). The intersite hopping
325
is t and the on-site repulsion U . In the standard method, one sets up basis
vectors for the Sz = 0 sector
|v1 >= |a a >, |v2 >= |a b >,
|v3 >= |b a >, |v4 >= |b b > .
One then looks for eigenstates (three singlets and one triplet)
=
4
i |vi >
(14.57)
i=1
U
t
H=
t
0
t t
00
00
t t
0
t
.
t
U
(14.58)
We can do with 2
2 (rather
bythe spin-disentangled
than
4 4)
matrices
10
00
0 t
1 2
, nb =
.
,K =
, na =
method with L =
00
01
3 4
t0
By (14.56),
a,b
a,b
(H[L]) | |
(14.59)
H| =
U 1 + t(2 + 3 )
t(1 + 4 )
.
t(1 + 4 )
U 4 + t(2 + 3 )
The reader can readily verify that this is the same as applying H in the form
(14.58) to the standard wave function (14.57) and then casting the result
in the form (14.51). Since we can apply H we can also diagonalize it. By
this device, we can work with matrices
whose dimensions is the square root
326
hF =r nr+
k nk +
{Vkr ck cr +h.c.},
k
Hv =0 d d,
(14.61)
(14.62)
where a linear coupling is assumes between the oscillator and the resonant
state:
Hev = g(d + d )nr .
(14.63)
The electron can be captured by the resonance and excite the plasmons, but if
eventually the resonance decays, the interaction with the plasmons is turned
o, otherwise a permanent shift of the oscillator coordinate may result. The
interaction can be very strong if the boson frequency resonates with dressed
electronic excitations. The purpose of the method is to calculate exactly the
the excitation amplitudes
v (E) = 0| dv cr
1
c+ |0
E H + i r
(14.64)
This approach should not be confused with other useful recursion methods,
which are completely dierent, despite the fact that they lead to continued fraction solutions. Among the most important examples, I recall the Mori generalized
Langevin Equation Method [208] and the Lee solution technique [209]; for a review
see Ref. [210]. However, quite recently, I became aware of a related solution of a
boson-boson model had been put forth previously by Sumi[100] developing a theory
of exciton polaritons.
4
For applications to many-photon eects, see Reference [78] and next Chapter.
327
1
1
1
1
=
+
Hev
EH + i Eh + i Eh + i
EH + i
yields
1
c |0 +
E h + i r
1
1
0| dv cr
Hev
c |0 =
E h + i
E H + i r
1
1
Hev
c |0
v0 G0rr (E) + 0| dv cr
E h + i
E H + i r
v (E) = 0| dv cr
(14.65)
where G0rr (E) is the retarded Greens function for g=0. In other terms, this
is the solution of the Fano problem,
G0rr (E) =< 0|cr
1
cr |0 >.
Eh + i
(14.66)
(14.67)
1
1
=
dv .
E h + i
E h v0 + i
(14.68)
Thus,
v (E) = v0 G0rr (E) + 0| cr
1
1
dv Hev
c |0 (14.69)
E h v0 + i
E H + i r
!
"
1
1
c gdv d + d cr
c |0 .
E h v0 + i r
E H + i r
(14.70)
Now note that, by reading the matrix element from right to left, the electron
is created, annihilated, created and annihilated again, thus the rst annihilation produces the electron vacuum; moreover, on the left of dv the only
boson operators are those contained in h. Therefore, the matrix element gets
factored.
= v0 G0rr (E) + 0| cr
1
c |0
v (E) = v0 G0rr (E) + g 0| cr
E h v0 + i r
1
0| dv d + d cr
cr |0 .
E H + i
Since
(14.71)
328
! "n
! "n1 ! + "n
d d+ = n d+
+ d
d,
(14.72)
0|d d = 0|d1
(14.73)
one gets
(14.74)
(E)
1
E
-1
v (E) = gv v (E)
(14.75)
(14.76)
G0 (E)
1
g2 G0 (E)G0 (E
0)
2g2 G0 (E0 )G0 (E20 )
3g2 G0 (E20 )G0 (E30 )
1
1.................................
(14.77)
from which, using the r.r. again, the other amplitudes with are easily derived. The interacting density of states (E) = Im()
gets satellites and
329
is deformed in a characteristic way (see Fig. 14.2). Note that G0rr (E) is not
a simple energy denominator as in Haydocks method and contains the full
band-theory information. This method has already been applied to a variety
of problems involving few electrons interacting with bosons, leading to exact
solutions in many cases. The present is a simple example, and more generally
the solution([124])[46]) is not given by a continued fraction.
is the sum of all the products of N elements in the matrix, one for each line
and column; each product is multiplied by the sign of the permutation Q
of the rst indices, when the product is ordered with the second indices in
ascending order 1 to N. Below Ill write simply A for Det A. Everybody knows
the Laplace development; choosing an arbitrary row index i,
()(i+j) aij Aij ,
(14.79)
A=
j
rc
where A is obtained from A by removing row r and column c. This is not the
only useful expansion, however. It often happens that the diagonal elements
are nite, that is, O(1), while the o-diagonal ones are O(V), where V is some
small parameter. In such cases, the Feenberg development is specially suited,
since besides being exact, it can serve as an expansion in powers of V.
Let Ai be the determinant obtained from A by removing the i-th row
and column; Aij the determinant of the matrix that one obtains from A by
removing rows and columns i and j; and so on. Chosen at will a diagonal
element aii , in each term of A it is multiplied by elements coming from rows
A
and columns
= i; thus a
is an antisymmetrized product of all the elements
ii
A
from other rows and columns, and this is just a
= Ai . Given a pair of
ii
A
indices i and j and the product pij = aij aji , pij is an antisymmetrized
A
= Aij ; one
product of the elements from other rows and columns, so p
ij
can continue in this way, ending with A123...N 1. Thus, one obtains a closed
development:
aij aji Aij +
aij ajk aki Aijk
A = aii Ai
j=i
k=j,i j=i
l=k,j,i k=j,i j=i
(14.80)
330
A convenient notation is
A = aii Ai
aij aji Aij +
aij ajk aki Aijk
j
j
k
(14.81)
where j means that the index j must be dierent from any other index
in the summand. Equating the Feenberg and the Laplace developments one
nds for i
= j; summarizing,
Aij = Ai , i = j
= ()(i+j) [aji Aij
ajk aki Aijk +
k
ajk akl ali Aijkl ], i
= j.
l
(14.82)
The sequences of dierent indices like those in ajk akl ali are called by
Swain[98] irreducible processes. To write down a general term of such an
expansion one can start from enumerating all the irreducible processes that
occur at a given order. The elements of the inverse matrix are:
(i+j)
(A1 )ij = ()
Aji
A
(14.83)
Ai
.
A
(14.84)
Let us introduce:
Di = AAi , Dij =
i
=
Dji = AAiji , Djk
A
A
Aij , Dijk = Aijk ,
A
Ai
il
il
Aijk , Djk = Aijkl ,
(14.85)
where where each D brings lower and possibly upper indices; the upper ones
are those of the A above, while the A below bears the upper and the lower
indices. Directly from the Feenberg development,
a a A
A
Di = AAi = aii j ij Ajii ij + j k aij ajk aki Aijk
i
Aijkl
l k j aij ajk akl ali Ai + ....
a a
a a aki
= aii j ijDiji + j k ijDjk
i
j
jk
aij ajk akl ali
l k j
+ ....
Di
(14.86)
jkl
and similarly
5
one uses the notation Gi since if A = z H these are resolvent matrix elements.
Dki =
331
(14.87)
a a
Diuv = aii j Dijiji + j k
a a ja a
l k j ij Djkikl li + ....
(14.89)
(14.90)
jkl
A
A Ai
=
= Di Dji = Dj Dij
Aij
Ai Aij
(14.91)
(14.92)
where
D1 = a11
D21 =
a12 a21
.
D21
(14.93)
(14.94)
332
where j=1 is forbidden, and so j=3; then, no k is acceptable and we are left
with
a23 a32
D21 = a22
.
(14.95)
D312
So,
G1 =
1
a11
a12 a21
a a
a22 231232
(14.96)
D
3
aii+1 ai+1i
,
Ci+1
Ci = Di1,2,...i1
(14.97)
(14.98)
(14.99)
()(i+j) Aji
j
bj
(14.100)
...
j Dij
jk
jkl
Di
Dijk
Dijkl
(14.101)
333
This case reduces to the previous one: provided r is such that the determinant
Ar does not vanish, one can assign an arbitrary value to xr and the reduced
system
aij xj = air xr , i = 1, ..., N, i
= r
(14.103)
j=r
with bi = air xr (excluding row and column r from all determinants) has a
unique solution
xi
air aij ajr aij ajk akr
= r +
+ ...
(14.104)
r
j Dr
jk
xr
Di
Dijk
ij
Problems
14.1. Show that if we can calculate the local Greens function, we can also
obtain any matrix element Gmn (E).
14.2. For the rectangular band model (14.20) n() =
continued fraction coecients.
(W 2 2 )
,
2W
nd the
14.3. For the semiellyptic band model, nd the continued fraction coecients.
14.4. The 33 linear system Ax = b with a non-singular matrix
11
11
1
1 a23 a21 1 1 a11 a13 1 1 a13
1
1
1
1
1
11
A =
1
11
11
detA
a33 a31 a31 a33 a23
1
11
11
1 a21 a22 1 1 a12 a11 1 1 a11
1
11
11
1 a31 a32 1 1 a32 a31 1 1 a21
1
a13 11
a23 1
1
a11 11
.
a21 1
1
a12 11
a22 1
2 04
2
d .
3
3c
(15.1)
d =
E cos(t);
(15.2)
the polarizability
is a second-rank tensor. If m is the pulsation of the
molecular vibrations,
=
0 +
1 cos(m t + )
(15.3)
where is an arbitrary phase. Thus, one expects elastic radiation at and inelastic components at m . The one at m and +m are called Stokes
and anti-Stokes radiation, respectively. Classically, their intensity should be
the same, actually the Stokes light is more intense, but this requires Quantum
Theory to explain.
336
Photon
We use the trasverse gauge, without scalar potential and vector potential
A = N
ei k r it ;
(15.4)
for the normalization we ask that a unit volume V contains a photon energy,
that is,
1
.
(15.5)
E E + B B d3 r = h
4 V
A ,
From E = t
B = rot A , one nds E E = B B = N 2 2 ; putting into
2
2
(15.5), 2N
and
4 V = h
A =
2
h
ei k r it ;
V
(15.6)
2
h
e
e d( A r )
dA
it
e
. Then, using
r
A r =
we
A
V
dt
dt
dt
(15.7)
Here,
2
h
(15.8)
d , VE = VA ,
V
describe photon absorption and emission, respectively.
VA = i
where VA and VE
337
t= |f
t= |f
1
t=t2
t=t1
t=t1
t= |i
t=t2
t= |i
a)
b)
Fig. 15.1. A photon scattering second order process: in a) the photon is absorbed
and then re-emitted, in b) it is re-emitted and then absorbed.
(15.9)
This would be the correct result if the times t1 and t2 were assigned by the
experiment. Since this is not the case, we must allow for the interference of
dierent values, by integrating over the intermediate times. A part Aif of
the amplitude comes from events in which t2 > t1 : the contribution to the
process is f |t where
|t =
dt2
t2
(15.10)
thus
a)
Aif (t) =
dt2
t2
(15.11)
338
a)
Aif (t) =
t2
dt2
dt1
(15.12)
dt2
i(HEi 1 )t1
t2
dt1
VA (1 )|i .
(15.13)
Thus,
a)
Aif (t)
=
i
VA (1 )|i
H Ei 1 i0
t
f |VE (2 )
i
VA (1 )|i (15.15)
H Ei 1 i0
and nally
i
Ef Ei + 2 1 i0
i
f |VE (2 )
VA (1 |i .
H Ei 1 i0
(15.16)
Ef Ei + 2 1 i0
i
f |VA (1 )
VE (2 |i ,
H Ei 2 i0
(15.17)
b)
(15.18)
339
Energy Conservation
The 1 photon is incoming, and must be switched on adiabatically; therefore
Aif (t) must have an innitesimal exponential growth. Letting x = Ef Ei +
2 1 ,
et
e2t
|Aif |2 2
, 0.
Aif
x i
x + 2
The transition probability per unit time is
2e2t
dPif
2
2(x), 0.
dt
x + 2
(15.19)
dPif
+ Ei Ef having k within d2 . The sum
dt on the modes with 2 = 1
(2 |k|), that is,
is the density of states d(2 )
k d2
V d2
V d2 2
d(2 ) =
.
(15.20)
d3 k(2 |k|) =
3
2)
(2)3 2
The incident photon is normalized to 1 in volume V ; the probability per unit
time to observe the outgoing photon is the dierential cross section
1
1
i
VA (1 )|i +
dif = 2 11f |VE (2 )
H Ei 1 i0
12
1
i
VE (2 )|i 11 d(2 ).
f |VA (1 )
(15.21)
H Ei 2 i0
Recalling (15.8),
2
dif = |M |
1 23 d2
,
h2 c4
(15.22)
where
i
1 d |i +
H Ei 1 i0
i
2 d |i .
f |
1 d
H Ei 2 i0
M = f |
2 d
(15.23)
340
1 df n
1 dni
2 dni
2 df n
(15.24)
M=
+
ni
1
nf
1
n
provided that no denominator vanishes (otherwise one speaks of resonant
Raman scattering which requires a separate treatment). Following Kramers
and Heisenberg, we may write
Rpq (2 )p (1 )q ,
(15.25)
M=
pq
Rpq
f |
d p |n n| d q |i f | d q |n n| d p |i
=
+
.
ni 1
nf 1
n
(15.26)
In the elastic f = i case, the tensor describes the Rayleigh scattering. Otherwise if 1 > 2 then Ef > Ei (Stokes transition); in the opposite case the
system gives energy to the photon (anti-Stokes transition).
Inversion Symmetry
We saw in Chapter 8 that in systems like the Benzene molecule possessing
inversion symmetry, Rpq = 0 unless |i and |f are of the same parity. Thus,
Raman-active modes are not seen in infrared absorption, and conversely infrared active modes yield no Raman scattering. infrared.
Low-frequency Scattering
For 1 0, the Raman tensor has a nite limit. Hence, the cross-section
(15.22) goes like 1 23 . The elastic scattering goes with 14 . This explains
why the sky is blue and the advantage of radio-astronomy over infrared or
optical observations of objects that are beyond interstellar clouds.
High-frequency Scattering
When 1 is large compared to the main absorption frequencies, we insert into
(15.26) the expansions
1
ni
1
2 + ,
ni 1
1
1
1
nf
1
2 + ,
nf + 1
1
1
and nd
(1)
(2)
Rpq = Rpq
+ Rpq
+
(15.27)
1
f | d p |n n| d q |i + f | d q |n n| d p |i = 0
1 n
(15.28)
1
ni f | d p |n n| d q |i + nf f | d q |n n| d p |i .
2
1 n
(15.29)
d
ni n| d q |i = n|[H, d q ] |i = in| dt
d q |i ,
d
nf f | d q |n = if | dt d q |n ,
(15.30)
with
(1)
=
Rpq
and
(2)
Rpq
=
341
Since
Rpq = 2
f | d p |n n| d q |i f | d q |n n| d p |i
1 n
dt
dt
d
i
d q d p |i .
= 2 f | d p d q
(15.31)
1
dt
dt
2
[ d p,
d q ] =
pq .
(15.32)
dt
m
Thus, for large 1 ,
N e2
Rpq =
if pq .
(15.33)
m12
No Raman scattering occurs. For the elastic scattering, remarkably the crosssection (15.22) becomes frequency independent:
1
12 4
2 2
1 N e2
1 1 d2
e
2
2
1
d = 11
(
)
(
)
=
N
|
2
1 | d2
pq
2 p
1 q1
m12
c4
mc2
2
= N 2 r2 |
2
1 | d2 ,
(15.34)
e
e2
mc2
where re =
1 and
2,
1 |2 = cos2 ().
|
2
(15.35)
342
|f
|f
|f
|f
3
2
1
t= |i
a)
|i
b)
|f
|f
|i
|i
d)
c)
|i
e)
|i
f)
(s, n, r, 1, 1 , 2 ) = fr
snr
(15.38)
Blowing up the sample is not the best way to study it; however the next Section
deals with interesting cases that cannot be treated at low order because the eld is
so strong that all the degrees of freedom are dressed and deeply aected by photons.
343
(s, n, r, 2, 1 , 2 ) = fr
(15.39)
(s, n, r, 3, 1 , 2 ) = fr
(15.40)
(s, n, r, 4, 1 , 2 ) = fr
(15.41)
(s, n, r, 5, 1 , 2 ) = fr
(15.42)
(15.43)
and
(s, n, r, 6, 1 , 2 ) = fr
(15.44)
This was my starting point for a calculation2 of SHG spectra from interfaces
[104], an unexplored subject at the time, in the European Esprit Project
EPIOPTICS. Experiments were planned to observe the SHG signal from
Si surfaces and interfaces. EPIOPTICS had the main motivation that SHG
may carry information from buried interfaces, without exposing them, which
is important for materials characterization. A photon is odd under inversion,
and it can be shown that to second order no SHG occurs from systems with
inversion symmetry such as Si. The surface breaks the inversion symmetry,
and with the available radiation sources a tiny fraction (even 1015 or less)
of the incident power was predicted and observed to go into the SHG signal
????.
With a similar formalism one can study other nonlinear eects like the
two-photon decay of the 2s level of the H atom. In the following, we shall
deal with many photon eects that occur when the elds are so strong that
low order diagrams cannot work.
In atomic and molecular problems, when no medium aects the eld propagation, my formulation is an alternative to a more traditional one for bulk
materials[105], which was based on the second-order susceptibility (2) .
344
with
e (k)
e
m| A
p |n =
A m|
p |n
mc
mc
where
p is the momentum operator. Therefore,
(k)
=
Mmn
k + b ),
Hk = A(k) L(b
k
(15.46)
(15.47)
operates on the electrons but not on the photons. The part of the
where L
Hamiltonian that describes the eld and its interaction is
H=
M /
0
k + b ) .
bk bk + A(k) L(b
k
(15.48)
the rotated vector with components k A(k) mk has the same length as the
original one,
(A(k) )2 .
||A|| =
k
(15.50)
then,
H=h
1 + d ).
dm dm + ||A||L(d
1
(15.51)
345
Mode Normalization
2
h
i
V e
2
hc2 2
h c2 V
=
(3) k,
V
V (2)3
(15.52)
(3)
k
where (3) k is the spread of the laser beam in k space. Actually,
(2)3 is
comparable to the inverse of the volume where the laser beam is coherent;
hence we can safely work with one eective mode if V is the coherence volume.
(15.53)
where He represents the electronic system, Hph the free photons, HI and
HI the interactions of the electrons with the laser and scattered photons,
respectively. The free eld may be described by the Hamiltonian
Hph = 0 d d + 0 dscatt dscatt ,
(15.54)
where d annihilates photons 0 from the laser while dscatt destroys scattered
photons. The interaction terms are:
(d + d )
HI = M
(15.55)
(dscatt + d ),
HI = M
scatt
(15.56)
and M
are operators acting on the electronic degrees of freedom.
where M
The laser photons belong to one mode and are in a coherent state such that
d|c = |c
(15.57)
s = d .
(15.58)
346
=H
e + H
ph + H
I + H ,
H
I
(15.59)
e = He + 0 2 + 2 M
(15.60)
ph = 0 s s + dscatt dscatt ,
H
0
(15.61)
I = 0 + M
(s + s ) = M
(s + s )
H
(15.62)
where
The initial state of the system is |g; 0 |g |0 , where |g stands for the
electronic ground state and |0 for the photon vacuum.
The scattered spectrum S(0 ) is proportional to the dscatt photon emission rate. Instead of deriving a general formal expression for te rate (see
Ref. [46]), here I consider a special case by a simple, physically motivated
treatment.
15.3.2 Dynamical Stark Eect
A photon impinging on a molecule or a solid can give raise to lots of processes,
but one can choose the symmetry of the system and the photon such that
just one transition is possible. Then the system-radiation interaction can be
understood by a two-level model. F Schuda, C R Stroud Jr and M Hercher
[106] using a dye laser and a beam of Na atoms realized this situation: the
o
F=3
3 2P 3
2
F=2
F=3
3 2P 3
2
F=2
2
3 S1
2
F=1
a)
F=2
F=2
3 S1
2
F=1
b)
of
Here, F is the total angular momentum resulting from the nuclear spin I =
21
11 N a
3
2
347
(15.63)
Here, m , am denote the unperturbed atomic levels and annihilation operators, nm = am am , g the electron-photon matrix elements, 0 the laser frequency. The scattered photons and their emission
H = dscatt dscatt + HI
HI = (a1 a2 + a2 a1 )g(dscatt + dscatt )
(15.64)
g
0
obtaining
Hph 0 s s + 0 (s + s ) + 0 2 ,
HI (a1 a2 + a2 a1 )g(s + s ) + 2g(a1 a2 + a2 a1 ).
We rearrange the terms of the transformed H0 as follows:
e + H
I + 0 s s,
0 = H
H
with
(15.65)
e = 1 n1 + 2 n2 + 0 2 + 2G a a2 + a a1 , G = g
H
1
2
(15.66)
!
!
"
"
I = 0 + g a a2 + a a1 s + s = M
s + s .
H
1
2
(15.67)
the initial state of the system is a1 |0 , where |0 is the vacuum for s and
dscatt photons. Note that g enters as G = g in Equation(15.66) and as g
in in equation (15.67). The two entries have quite dierent consequences.
4
For simplicity, we ignore the MF multiplicity and model this as a two-level
system for a spin-less eective particle.
348
h 0 .
(g)2 g 2 n
a0
V
2
(15.69)
(15.70)
(15.72)
Since G
2 1 , states are little aected by the tildes, and the matrix
element between almost degenerate states is6
1, n+1|g(a1 a2 +a2 a1 )(d+d 2)|
2, n gn+1|d |n = g n + 1. (15.73)
5
|2
0
|1
|
1, 3
|
2, 2
|(2)
|
1, 2
|
2, 1
|(1)
|
1, 1
|
2, 0
|(0)
|
1, 0
a)
349
|
1, 0
b)
c)
Fig. 15.4. Resonant radiation interacting with a two-level system. The energy
spectrum is represented a) according to a simple one-photon view b) in the manyphoton description, involving an innite ladder of states, neglecting the g(a1 a2 +
a2 a1 )(s + s ) term c) with the g(a1 a2 + a2 a1 )(s + s ) term included.
2,n
|1, 0 ground state and an innity of close doublets; the doublet |1,n+1|
2
n + 1 |2,
n and the splitting is 2g n + 1.
arises from |1,
Scattered Light
Up to now we have dressed the atomic levels with photons from the laser obtaining an innite double ladder of states, but the scattered photons still have
to be considered. The scattered light arises from the spontaneous emission
between the dressed states, according to A. Einsteins A coecient
Amn =
3
4mn
d2 ,
3hc3 mn
(15.74)
+ (n) (n 1) 0 + g( n +
1
+
n) 0 + 2g n
+ (n) + (n 1)
0 + g( n + 1 n) 0
(n) (n 1)
0 +
g( n + 1 + n) 0
(n) + (n 1) 0 + g( n + 1 + n 0 2g n)
(15.75)
350
|+ (n)
n0
| (n)
|+ (n 1)
(n 1)0
| (n 1)
a)
b)
Fig. 15.5. The contribution of two consecutive doublets to the scattered radiation.
a) the spontaneous emission lines b) the resulting spectral intensity versus .
6
ea an 5
( 0 2g n) + ( 0 + 2g n)
4 n=0 n!
| 0 |ea an
(( 0 )2 4g 2 n).
2
n!
n=0
(15.76)
To obtain the shape of the side peaks, one can rewrite Equation (6.33) in
the form:
(zaz0 )2
n
2g2
a
e
(z nz0 )
, a 1,
(15.77)
ea
n!
z0 2a
n
with z = ( )2 and z0 = 4g 2 . Finally, the spectrum may be approximated
by
(x2 a)2
1
|x|
exp[
]+
.
(15.78)
S(x) =
2
2a
2 + (2x)2
4g 2a
351
S(x)
0.15
-10
10
Fig. 15.6. The characteristic shape of S(x) versus x for a = 100 and = g.
Problems
15.1. Prove Equation (13.136).
Part V
16 Quantum Phases
2
(p eA
c )
+ eV (x)
2m
(16.1)
where p is the kinetic momentum and A the vector potential. Both are unobservable; one could have started with new potentials A , V such that
A = A + , V = V
1
c t
(16.2)
e
p (p ) , i
(ih +
) .
c
t
t c t
(16.4)
One nds that the new wave function is related to the old by
(x, t) = (x, t) exp[
ie(x, t)
],
hc
(16.5)
that is, substitution into (16.3) gives the correct equation for . The gauge
changes the phase at any point and at any time in an arbitrary way. The
matrix elements of the coordinates are trivially invariant, while those of the
canonical momentum are shifted by pmn (p + ec )mn ; the Schrodinger
theory is invariant because what matters is the mechanical momentum p ec A
which remains unshifted. Thus, the theory is gauge invariant. Equation (16.5)
also holds in Klein-Gordon and Dirac theories.
356
16 Quantum Phases
(16.6)
cos() sin() 0
(16.7)
{aik } = sin() cos() 0
0
0 1
and x appears like x rotated clockwise. How does (16.6 ) transform to K ?
Since the Pauli equations must hold in the rotated system,
( B ) = ,
(16.8)
Ri R = aik k
(16.9)
iz
2
2 2
h
(x, t) + U (x, t) (x, t) = ih (x, t).
2
2m x
t
(16.10)
We go from the reference K of coordinates (x, t) to a reference K of coordinates (x , t) moving with speed v, via the Galilei transformation x =
x vt, t = t. Galilean covariance requires that the problem in K becomes
2 2
h
(16.11)
357
t
x
= v x
(16.13)
2 2
h
(16.14)
(16.15)
mvx mv 2 t
.
h
2h
(16.16)
358
16 Quantum Phases
(16.17)
2i
0
b
A d
r
a
(16.18)
where
hc
= 4 107 Gauss cm2
(16.19)
e
is the ux quantum or uxon. In the case of H2 this can be gauged away,
but with three or more atoms the physical meaning is that a magnetic ux
is concatenated with the molecule; changing by a uxon has no physical
meaning, however.
0 =
(16.20)
359
the system at any time would have3 . To be specic, let the motion start at
time 0 and end at time T ; therefore the duration T must be long enough to
meet the adiabatic assumptions. We wish to solve
i
h
n = H[R(t)]n
t
(16.21)
(16.22)
We shall nd that the integral of (16.21) bears a real Berry phase [65] n (t),
which turns out to be due to the topology of H in parameter space. This phase
can then be dropped or altered by a gauge transformation at any t, but a
net phase acquired in a closed path cannot be gauged away. To calculate the
phase, lets expand on the instantaneous basis:
n (t) = an [R(t)]cn (t) +
cm am [R(t)], cn (0) = 1.
(16.23)
m=n
By the Kato adiabatic theorem 4 under our assumptions (nondegenerate discrete spectrum and slow enough evolution) only the rst term remains: the
system does not move from the n-th eigenstate. The adiabatic solution has
|cn (t)| 1, and the only eect of the evolution is on the phase, namely,
cn (t) = ei(n (t)+n (t)) ;
here
hn (t) =
(16.24)
dt En [R(t )]
(16.25)
n (t) = ( En + i n + R R )an e h
t
h
+
cm am +
cm a m .
m=n
t
0
(16.26)
m=n
We drop the last term which is negligible, because the coecients of the
m
= n terms are small and the derivative is also small: these are second
order terms. Using (16.24), 0 = (i
h t
H)n becomes
3
The original paper by Berry is based on the adiabatic hypothesis which makes
the analysis easier, however the existence of the Berry phase does not depend on
this assumption.
4
The proof is clearly discussed in Ref.[194].
360
16 Quantum Phases
t
i
0 =
h n + ih R R an e h 0 [ n
[cm Em ]am
+
m=n
(16.27)
The matrix element looks similar to a momentum average, but the gradient
is in parameter space. The overall phase change is a line integral
T
(16.28)
n (C) = i an |R |an d R
C
()|( + )
|()|( + ) |
(16.29)
N
i,i+1
i=1
gauge invariant. This is an open path Berry phase; |(1 )|U |(1 ) | = 1
because the spectrum is not degenerate. In the case of a symmetry connecting
1 to itself, we speak of a single-point Berry phase ; by (16.29),
ei = |U | = arg(|U | ).
(16.30)
361
n (C) =
An dR
(16.31)
n (C) =
Bn
rot A n
n dS
n dS.
(16.32)
S
B n = ImR an | |aR an = Im
R an |am am |aR an .
m
Note that
an |R an [R]
is imaginary, (the terms with m
= n have arbitrary phases). Thus under the
Im operator we may remove the real m=n term. If you think that saving one
term is not an achievement, wait a moment. We get:
362
16 Quantum Phases
am |R H|an
.
Em En
(16.34)
B adiabatically, sweeping a solid angle (C). Berry has shown that the
geometric phase is n (C) = n(C).
16.1.2 Polarization of Solids
The modern theory of polarization in solids is based on these ideas; for a
review, see [16]. Consider a periodic solid with Hamiltonian
1 2
p + V.
2m i=1 i
N
H=
(16.35)
The usual denition of the dipole moment of a system in terms of the density
(r), which applies to molecules, is:
d = e d
r
r (
r ) = e0 | R |0
(16.36)
where |0 is the ground state and
eR = e
r i,
N
(16.37)
i=1
summed over the N electrons,is the dipole operator. There are problems in
applying this to solids because: 1) in solids we should like to use periodic
boundary conditions, and the dipole operator is forbidden, since it takes
outside the Hilbert space of periodic functions 2) we should like to deal with
a bulk property, while (16.36) depends crucially on the surfaces 3) one never
measures anything like (16.36). What one can do is rather to measure the
d
(e R ) ) coming out from the crystal under
current (which is proportional to dt
some time-dependent probe, like a mechanical stress; we suppose that the
experiment is done adiabatically.
363
Adiabatic Current
One can dene the current density j = T rj,
j=
N
P
,
P
=
pi
mL3
i=1
(16.38)
in terms of the total momentum P , the electron mass m and the super-cell
size L; the polarization change in the experiment is
P = dtj(t) .
(16.39)
The density matrix (t) in the adiabatic limit was given by Niu and Thouless
[80] in terms of the eigenstates |n, t ; the instantaneous density matrix
i = |0, t 0, t|
(16.40)
(16.41)
d
[i + ] = [H, ];
dt
(16.43)
d
in the adiabatic case dt
is negligible (time derivatives are small, the correction is small). Hence, we are left with
d
i = [H, ].
dt
(16.44)
d
d
d
d
0, t|i |n, t = 0, t| i |n, t + 0, t|i |n, t +0, t|i | n, t ; (16.45)
dt
dt
dt
dt
d
d
but dt
0, t|i |n, t = 0 also vanishes (since dt
0, t|0, t = 0) and
0, t|
d
d
d
d
i |n, t = 0, t|i | n, t = 0, t| n, t = 0, t|n, t ,
dt
dt
dt
dt
(16.46)
364
16 Quantum Phases
where the last equality follows from 0, t|n, t = 0. Taking matrix elements of
(16.44) we obtain the desired correction
0n = i
t|n, t
0,
.
E0 En
(16.47)
Taking into account Equations (16.38 ) and (16.42) the average current is
j =
ie
h 0 |n n |P |0
+ c.c.
mL3 n>0
En E0
(16.48)
Idealized Experiment
Rather than a mechanical stress, we can more simply (from the theorists
point of view) use ingenuity to introduce a tiny electric eld without disturbing the periodicity of the solid. Consider modifying the Hamiltonian (16.35)
to read
N
1
(pi
hk)2 + V.
(16.49)
H=
2m i=1
This device [81] corresponds to a constant vector potential (that will vary
slowly in time) and would represent a gauge change, were it not for the
periodic boundary conditions; running through the supercell of length L an
electron collects a magnetic ux. Indeed, this is called Hamiltonian with a
ux. In order to make this compatible with the periodicity, we must choose
()
with components
k appropriately; we introduce vectors k
()
2
;
L
(16.50)
()
U (k) = ei k R ,
(16.51)
()
U ( k ): picking k = k , which makes U allowed, we satisfy the modied
Schr
odinger equation and the periodic boundary conditions by writing the
ground state of (16.49)
()
()
|0 ( k ) = U ( k )|0
(16.52)
()
z () = 0 |U ( k )|0 = |z () |ei
365
(16.53)
()
0 |U ( k ) |k k |U ( k )|0 ;
|z () |2
k
|n n | h kP |0 |
m
.
E
E0
n
n>0
(16.54)
Normally in writing this rst-order expansion one does not care about the
phase, but here we must keep track of (16.53); so we write using (16.50)
|0 (k) = ei {|0 +
2
h |n n |P |0 |
}.
mL n>0
En E0
(16.55)
(16.56)
2
h 0 |n n |P |0 |
}.
mL n>0
En E0
(16.57)
0 |U |0 + 0 |U | 0
]
0 |U |0
(16.58)
e
.
2L2
(16.59)
366
16 Quantum Phases
The connection with the Berry phase is established, and the polarization
change P is proportional to .
The actual calculation can be accomplished by expanding in determinants.
Indeed, if |0 = Det|1 N | is a determinant,
U (k () ) |0 = exp ik
ri Det|1 N |
i
also is, with spin-orbital i multiplied by exp [ik ri ] . The overlap (16.53) is
computed as a determinant of one-body overlaps.
Problems
16.1. Given the two reference systems of Section 16.0.5, if in K
k2 t
(x, t) = eikxih 2m
Long ago, Kohn and Luttinger [174] proposed that a superconducting instability should occur in Jellium. They focussed on pairs of parallel spin electrons
of large relative angular momentum L. The large L and the triplet state keep
the electrons apart, so the Coulomb repulsion is particularly mild in such
pairs. At long distances, the screened interaction undergoes Friedel oscillations due to the singularity of the dielectric function at 2kF (See Section
12.1.1). This means a slight over-screening of the repulsion in some distance
ranges: one gets an eective attraction from the repulsion via a quantum
mechanical correlation eect. Attraction in some distance ranges does not
necessarily imply binding, but Kohn and Luttinger suggested that this effect could indeed produce pairing. While no such superconducting instability
appears to be relevant to ordinary metals, the paradoxical theoretical idea
that attraction could result from repulsion is fascinating and the discovery of
high-Tc superconductivity [175] has stimulated a hot discussion on the possibility that something similar is realized in the Cuprates (although the pairs
are singlets with L=2, so some important modication is needed). The fact
that TC can be above liquid N2 (rather than liquid He) temperatures, and the
evidence that this occurs in strongly correlated materials suggests to part of
the community that the mechanism must be dierent from the conventional
phonon-assisted BCS one[181]. Other Authors prefer a conventional, or an
enhanced[182] phonon-assisted mechanism. New superconductors including
doped Fullerenes and Carbon nanotubes[197] complicate the riddle.
Anyhow, strong correlations and phonons are there, so they are likely to
be involved, with a complex trade between the various degrees of freedom;
some modeling is in order. Then, it looks natural to start by the repulsive
trivial Hubbard model (1.64), or some of its variants, in 2 dimensions, while
keeping in mind the claims that such a model cannot superconduct at all [168]
[169] or at least cannot have superconducting long-range order [170]. An effective attractive force comes from second-order eects, e.g., the exchange of
spin uctuations, but it must overcome the strong direct Hubbard repulsion
to give pairing. Bickers and co-workers[162] explored the consequences of a
spin density wave instability within the RPA approximation using the trivial
Hubbard Hamiltonian. They found a spin density wave phase and a superconducting phase with pair-wavefunctions of dx2 y2 symmetry. These ndings
368
were conrmed by using the FLEX approximation[163] which treats the uctuations in the magnetic, density and pairing channels in a self-consistent
and conserving way. Renormalization Group (RG) methods[164] are a well
controlled alternative approach to deal with Fermi systems having competing singularities. The RG has been used by several authors[165],[166],[167] to
study the coupling ows at dierent particle densities. In agreement with the
previous ndings, RG calculations show a d-wave superconducting instability away from half lling driven by the exchange of spin- or charge-density
uctuations.
States of dierent symmetry may be separated by tiny energies per electron, so it is very desirable to have exact data to rely on and observe pairing
or get clear-cut results. The 1d Hubbard model is solved[73] by Bethe Ansatz
techniques (Chapter 18) but does not show pairing. In 2d one can compute
the exact ground state of small clusters by Lanczos techniques (Section 14.1).
A pairing criterion that we shall discuss below was given by Richardson in
the context of nuclear physics[176]. Dening
(N
+ 2) = E(N + 2) + E(N ) 2E(N + 1)
(17.1)
(N
+ 2) < 0
(17.2)
Fig. 17.1. Pictorial representation of Equation (17.1);the large circles stand for
the interacting N-body system, the black circles represents the added particle, and
the wiggly line is the interaction, which remains after the simplication.
build repulsive models that yield pairing and superconducting ux quantization, and that the two properties have a strong mathematical connection to
each other. Both require special symmetric clusters ( it was known that pairing occurs in the 4 4 system[177] but the simpler examples like CuO4 and
the general case had not been discovered yet) and derived an analytic theory of pairing interactions (W=0 pairing). We reported several detailed case
studies where the formulas are validated by comparison with the numerical
data. The results support general qualitative criteria for pairing induced by
the on-site repulsion only. Our approach successfully predicts the formation
of bound pairs, and explains why other ingredients like strong o-site interactions are needed to force pairing in other geometries. The analytic approach
369
is also needed to predict what happens for large systems and in the thermodynamic limit. Understanding the implications of the small cluster results is
not trivial since the computed pair binding energies show a rapid decrease
with increasing the cluster size; so several authors on the basis of the numerical data on 44 or even 66 clusters wrongly concluded that pairing in the
Hubbard model as a size eect. However, much larger cells (at least 3030)
are needed to estimate the asymptotic behaviour and the analytic treatment
suggests that pairing with a reduced but substantial binding energy persists
in the full plane. Details may be fond in a recent review paper[173].
17.0.3 W = 0 Pairing in Cu-O Clusters
In this Section, we illustrate the concept of W = 0 pairs and the way they
become bound states, by using examples with a geometry relevant for the
Cuprates. Nevertheless, the simple square Hubbard plane and its symmetric
clusters like the 4 4 cluster with periodic boundary conditions also present
W=0 pairing[189]. Moreover, Hubbard models with the geometry of Carbon
CuO4
Cu5 O4
Cu5 O16
Fig. 17.2. Allowed clusters that were used to show pairing and superconducting
ux quantization by exact diagonalization methods in Ref. [172]; the holes were 4,
and the number of congurations involved in the case of Cu5 O16 is 44,100.
Nanotubes (CNT) [188] allow W=0 pairing away from half lling, opening the
possibility that even there the mechanism of superconductivity is basically
electronic [207]. Here we focus on the three-band Hubbard Hamiltonian
H = K + W + Wosite
(17.3)
where introducing the site energies p and d refering to Oxygen and Copper,
the kinetic term is
370
K=t
jj ,
ij,
pj pj + d
ni + p
i,
nj ; (17.4)
j,
here ts are hopping parameters; pj (di ) destroy holes at the Oxygen (Copper)
ions labeled j (i) and n are the number operators; ij refers to pairs of nearest
neighbors. The interaction terms are
W = Ud
ni ni +Up
nj nj ,
Wosite = Upd
ni nj . (17.5)
i
ij,
For large enough clusters, the local physics is well described by an innite plane,
and the cluster shape is no longer important, but such large systems are outside the
scope of diagonalization methods. Yet, our approach emphasizes the importance of
the symmetry in the mechanism and predicts that any important distortion of the
lattice readily destroys pairing.
371
square Cu-O cluster, CuO4 , see Fig. 17.2, is also the simplest2 where W = 0
pairing occurs, and we shall use it as a prototype. The one-body levels and
their symmetry labels are reported in Table 17.1. Let us build a 4-hole state
A
Ex Ey B1
1
A1
2
4 + 0 0 2 + 4 + 2
Table 17.1. One-body levels of the CuO4 cluster in units of t as a function of the
dimensionless parameter tpp /t.
in CuO4 (see Fig. 17.3 b)). The rst two holes go into a bonding level of A1
symmetry; this is a totally symmetric (1 A1 ) pair. For negative , the other
two holes go into a non-bonding level of E(x, y) symmetry, which contains 4
spin-orbital states. Labelling sites as in Figure
17.3 a),
the creation
operators
p3
p1
4
Using the E orbitals according to the Pauli principle one obtains
=6
2
dierent pair-states. The irrep multiplication table allows
for labeling
them
?
?
?
A2 B1 B2 . Proaccording with their space symmetry: E E = A1
jecting according to usual rules we see that A2 is a spin-triplet, 3 A2 , while
the remaining irreps are spin-singlets, 1 A1 , 1 B1 and 1 B2 . One can readily
verify that the B2 singlet operator
1
bB 2 = Ex,
(17.6)
Ey,
+ Ey,
Ex,
2
=
for the degenerate orbitals are Ey
1
2
p 2 p4
=
Ex
1
2
(4)
has a minimum at U 5 t and it is negative when 0 < U < 34.77 t.
2
372
(4)
2
B1 (x2 y 2 )
0
E(x, y)
U
t
.01
12
.02
.03
4
a)
A1
b)
.04
c)
Fig. 17.3. a) The CuO4 unit. b) One-particle energy levels for < 0 with their
symmetry labels and their llings in the 4-particle ground state according to the
aufbau principle. Two fermions belong to the doubly degenerate E(x, y) representation. For 0, the Group becomes S4 C4v and the B1 (x2 y 2 ) level merges
with E(x, y) giving raise to triple degeneracy. Note that among the irreps of C4v
(see the Character Table in Appendix II) A2 and B2 are not represented in this
scheme. Below we shall see that this observation is crucial for W=0 pairing theory.
c)(4)
versus Ut for this cluster is negative in a wide range (it becomes positive
U
for t > 34.77 ) and the maximum pairing occurs at U 5t; this is not a weak
coupling phenomenon.
(4)
= 0.53 eV). The dependence of (4)
on the input parameters has been
studied in detail; also, we have found that any lowering of the symmetry is
373
Fig. 17.4. The second-order exchange interaction that at weak coupling can be the
leading contribution to the eective attraction, depending on the lattice and on the
parameters (see Ref. [172]).
interaction can be found by a canonical transformation (basically a SchrieerWol-type argument) which is long and not easily summarized, so I defer the
reader to the original paper [196]. The results reduce to those of perturbation
theory in the weak coupling case. Moreover, the analytical results compared
semi-quantitatively with those from exact diagonalization of several clusters
at various llings including the 4 4 Hubbard cluster[190]. Being predictive
in strongly correlated problems is the exception, rather than the rule, but
in this case simplicity has emerged from symmetry. Applying the formulas
to supercells of increasing size, it was possible to extrapolate the results to
the thermodynamic limit, showing that < 0 is not a size eect and could
well contribute a pair binding energy in the 10 meV range. The inclusion of
phonon eects in the W=0 scheme was envisaged in Ref. [70].
17.0.5 The W = 0 theorem
A general theorem puts useful restrictions on the many-body ground state
symmetry. Consider a Hubbard model4 H = H0 + W on a lattice , where
tij cj ci
(17.7)
H0 =
ij,
4
374
Ui ni ni ,
(17.8)
/ E W P () | = 0.
(17.9)
Then we have
ni ni | () = () (i, i)ci ci |0 () (i, i)|i , i .
Consider the projection operator P () on the irrep : since
() (i, i)|i , i =
() (i, i)|i , i ,
P ()
i
375
where |i
i |0 . Now it is always possible to write |i in the form
= c()
|i =
c
(i)| where | is the one-body eigenstate of H with
E
spin belonging to the irrep . Hence, if
/ E, P ( ) |i = 0 and so
P ( ) |i , i = 0. Therefore, if | () is a two-hole singlet eigenstate of the
kinetic term and
/ E, then it is also an eigenstate of W with vanishing
eigenvalue, that means a W = 0 pair.
Q.E.D.
Remark : Suppose we perform a gauge change in H such that G0 is preserved; clearly, a W = 0 pair goes to another W = 0 pair. Thus, the theorem
implies a distinction between symmetry types which is gauge-independent.
Remark :The complete characterization of the symmetry of W = 0 pairs
requires the knowledge of G0 . A partial use of the theorem is possible if
one does not know G0 but knows a Subgroup G0< . It is then still granted
that any pair belonging to an irrep of G0< not represented in the spectrum
has the W = 0 property. On the other hand, accidental degeneracies occur
with a Subgroup of G0 , and by mixing degenerate pairs belonging to irreps
represented in the spectrum one can nd W = 0 pairs also there.
17.0.6 Examples
CuO4
The CuO4 cluster is a simple example of the W=0 theorem. If
= 0,G0 = C4v ;
of the W = 0 pair belongs to the irrep B2 which is not represented in the
spectrum (see Table 17.1). A2 is also not represented, but projecting any twobody state one nds nothing. If = 0, a three-times degenerate one-body
level exists due to an accidental degeneracy between E(x, y) and B1 orbitals;
with 4 interacting fermions, pairing occurs in two ways, as A1 and B2 are
both degenerate ground states. The extra degeneracy cannot be explained
in terms of C4v , whose irreps have dimension 2 at most. For = 0 any
permutation of the four Oxygen sites is actually a symmetry and therefore
G0 is enlarged to S4 (the group of permutations of four objects). S4 has the
irreducible representations A1 (total-symmetric), B2 (total-antisymmetric),
E (self-dual), T1 and its dual T2 , of dimensions 1, 1, 2, 3 and 3, respectively.
These irreps break in C4v as follows
A1 = A1 , T1 = B1 E, T2 = A2 E, B2 = B2 , E = A1 B2 .
Labelling the one-body levels with the irreps of S4 , one nds that E is not
contained in the spectrum and thus yields W = 0 pairs:
1
E =
A1
B2 .
(17.10)
376
= B1
B1 + Ex
Ex + Ey
Ey
bA
1
3
3
(17.11)
for the total-symmetric pair and Eq. (17.6) for the B2 component.
The 4 4 Hubbard cluster with periodic boundary conditions
Table 17.2. One-body spectrum for t = 1. The irrep symbols will be explained
shortly.
1 2 3 4
5 1 4 8
5 6 7 8
6 2 3 7
d:
9 10 11 12
10 14 15 11
13 14 15 16
9 13 16 12
(17.12)
377
C1
C2
C3
C4
C5
C6
C7
C8
C9
C10
I
t22 C4 [2] x C2
C2 d C2 t22 d C2 [1] C43 t02 d
C11
C12
C13 C14 C15
C16
C17
C18
C19
C20
C43 t20 d C2 t01 C2 x [1] C4 C2 t01 d C2 t12 d C2 x [1]d C4 [1]d C2 [1]d C4 [1]
Table . Here, we report one operation for each of the 20 classes Ci ; the others can
be obtained by conjugation. The operations are: the identity I, the translation tmn
of m steps along x and n along y axis; the other operations C2 , C4 , , are those of
the Group of the square and are referenced to the centre; however,C2 [i], C4 [i], [i],
and [i] are centered on site i.
The complete Character Table of G is shown below.
G
A1
A1
B2
2
B
1
2
1
2
3
4
1
2
3
4
1
2
3
4
1
2
C1
1
1
1
1
2
2
3
3
3
3
4
4
4
4
6
6
6
6
8
8
C2
1
1
1
1
2
2
3
3
3
3
-4
-4
-4
-4
6
6
6
6
-8
-8
C3
1
1
-1
-1
-2
2
3
3
-3
-3
-2
-2
2
2
0
0
0
0
-4
4
C4
1
1
-1
-1
-2
2
3
3
-3
-3
2
2
-2
-2
0
0
0
0
4
-4
C5
1
1
1
1
2
2
3
3
3
3
0
0
0
0
-2
-2
-2
-2
0
0
C6
1
1
1
1
2
2
-1
-1
-1
-1
0
0
0
0
-2
-2
2
2
0
0
C7
1
-1
1
-1
0
0
-1
1
-1
1
-2
2
-2
2
-2
2
0
0
0
0
C8
1
-1
1
-1
0
0
-1
1
-1
1
2
-2
2
-2
-2
2
0
0
0
0
C9
1
1
1
1
2
2
-1
-1
-1
-1
0
0
0
0
2
2
-2
-2
0
0
C10
1
-1
-1
1
0
0
-1
1
1
-1
-2
2
2
-2
0
0
-2
2
0
0
C11
1
-1
-1
1
0
0
-1
1
1
-1
2
-2
-2
2
0
0
-2
2
0
0
C12
1
-1
1
-1
0
0
-1
1
-1
1
0
0
0
0
2
-2
0
0
0
0
C13
1
-1
-1
1
0
0
-1
1
1
-1
0
0
0
0
0
0
2
-2
0
0
C14
1
1
-1
-1
-2
2
-1
-1
1
1
0
0
0
0
0
0
0
0
0
0
C15
1
1
1
1
-1
-1
0
0
0
0
-1
-1
-1
-1
0
0
0
0
1
1
C16
1
1
1
1
-1
-1
0
0
0
0
1
1
1
1
0
0
0
0
-1
-1
C17
1
1
-1
-1
1
-1
0
0
0
0
-1
-1
1
1
0
0
0
0
1
-1
C18
1
1
-1
-1
1
-1
0
0
0
0
1
1
-1
-1
0
0
0
0
-1
1
C19
1
-1
1
-1
0
0
1
-1
1
-1
0
0
0
0
0
0
0
0
0
0
C20
1
-1
-1
1
0
0
1
-1
-1
1
0
0
0
0
0
0
0
0
0
0
378
tab tab e 0
Adr
In small clusters, W=0 pairing requires full C4v symmetry; moreover, the ux
tubes must be inserted in such a way that the symmetry is not distorted. We
must arrange the thought experiment in such a way that the perturbation
due to the eld is gentle, and does not destroy the pairing. In small clusters,
that we may stud more easily, the current carrying bonds must be so weak
that the current screening the eld is not excessive and the change in the
or less. When this is
ground state energy due to the eld is of the order
done, the SFQ already obtains with a few pairs and even one pair yields a
well dened double-minimum pattern. This means that although in textbooks
SFQ is explained in terms of the order parameter, it is basically a single-pair
symmetry property. Going to the thermodynamic limit is necessary to achieve
macroscopically large barriers that localize the system in one of the minima.
The simplest case is CuO4 and the insertion of ux tubes is shown in
Figure 17.4 a). In order to preserve the symmetry, a ux is inserted in each
of the four Cu-O-O triangles or plaquettes of the cluster. The dotted bonds
must represent negative hopping integrals tpp of the order of 50 meV or less,
otherwise the pairing mechanism breaks down and paramagnetism prevails.
Figure 17.4 b) shows the ground state energy of the cluster versus . The
ux dependence was computed with p d = 3.5, t = 1.3, Ud = 5.3, Up = 6.
This pattern is characteristic of SFQ. The system is diamagnetic at small
379
a)
E4 () (eV)
Ud = 5.3, Up = 6 eV
Ud = Up = 0
E4 () (eV)
8.006
5.15
1
4
1
2
3
4
7.986
5.05
7.966
1
4
1
2
b)
3
4
1
c)
Fig. 17.6. Rings of clusters threaded by ux tubes allow to study the propagation
of bound W=0 pairs. The dotted line represents hopping between two consecutive
clusters that must be designed in such a way that the square symmetry of each
cluster is preserved.
380
(its energy increases with ux) but reaches a maximum due to a level crossing
and goes trough a second minimum at = 20 ; = 0 is equivalent to = 0.
The two ground states at = 0 and = 20 are of comparable depth and are
due to W=0 pairs of dierent symmetry.
Fig. 17.7. Inserting 4 ux tubes, each carrying a ux in a CuO4 unit in the plane,
the magnetic ux through any path in the plane must equal times the number
of tubes encircled by the path. This is achieved by choosing the vector potential
according to the above pattern, with A dr = 0 along each bond without arrow,
A dr = 2 along each oriented bond with the integration path parallel to the
arrow. For a generic ux in each tube any reection of C4v reverses the arrows
( ) and thus C4v is broken and only the commutative Z4 Group of the
rotations survives. W=0 pairs are forbidden in the eld and pairing is destroyed.
However, at = 20 , one nds A dr = 40 and the Peierls prescription is t it
on any bond with arrow; now any reection implies which is just a gauge
change, replacing it by it. This is equivalent to a unitary transformation S that
changes the signs to all the Cu orbitals along the diagonals except the central one.
Since S reverses the arrows, S is a symmetry.
Indeed, as we saw earlier, two W=0 irreps exist for tpp = 0. SFQ is a
correlation eect and disappears for Ud = Up = 0; without correlation the
cluster shows a normal, paramagnetic behaviour, shown in Figure 17.4 c)
Similar phenomena have been observed in exact diagonalization work with
larger clusters (up to 21 atoms) in Ref. [172] and with rings of clusters. Larger
systems can be obtained by joining several clusters in such a way that bound
pairs can propagate. Rings of clusters threaded by ux tubes also show SFQ,
and Figure 17.5) illustrates the general idea of how such systems can be
381
built,in such a way that the square symmetry of each cluster is preserved.
The SFQ pattern is quite similar[18],[?] to the one discussed above; this time
the inter-cluster hopping must be a weak link in order to make the magnetic
perturbation gentle enough. By projecting on the low energy sector, we can
solve very large clusters in an essentially analytic way and show that the
superconducting behavior remains independent of size.
This suggests that the SFQ pattern must be a property of the Hubbard
model which extends to the thermodynamic limit and is essentially related
to the symmetry. The minimum at = 0 is clearly due to the W=0 pair,
yielding a nondegenerate ground state (when there s degeneracy, the ground
state must be paramagnetic, since the system interacts with the eld by a
Zeeman mixing of the degenerate components and gains energy). The eld
tends to destroy the pairs and does so by lowering the symmetry; indeed the
reections are broken. The second minimum at = 20 is due to the lattice
which for this particular ux recovers the ux symmetry. This is actually a
theorem, and the proof was obtained by a construction shown in gure 17.6
. One can see how one can choose the gauge for the vector potential when the
4 ux tubes are inserted like in gure 17.4 in a CuO4 unit belonging to a full
plane. From the construction in the Figure one sees that a general ux in the
ux tubes destroys the W=0 pairing by breaking the symmetry (C4v Z4 )
but at = 20 the full symmetry is restored, because the reections are
equivalent to a gauge transformation. The ux quantization pattern arises
from a level crossing between the ground states at = 0 and = 20 .
Models systems capable of superconducting ux quantization such as rings
of CuO4 clusters can also be operated[192] as Josephson junctions. One needs
to include a barrier and introduce a mild violation of the symmmetry to
produce avoided crossings; Josephson and inverse-Josephson (Shapiro) eects
then appear as adiabatic responses to slowly changing ux (t).
Recently, exact studies on small Hubbard clusters at nite temperatures
by Fernando et al.[193] have given useful insight on phase separation and
pairing.
Problems
17.1. Find how the irreps of the optimal Group G split in C4v .
17.2. Using the W=0 theorem and the Optimal Group, nd the irreps with
no double occupation in the 4 4 lattice .
18 Algebraic Methods
n.n.
x,y
cx cy + U
nx nx ;
U > 0.
(18.1)
1
(n (x) n (x)) , S + =
c (x)c (x),
2 x
x
(18.2)
(18.3)
in the site representation has S=0 and is singlet. A two-site system is the
most trivial bipartite lattice; with two electrons it becomes the model system
that we met in Sect. 14.2; there, I used it for illustrating the form (14.51),
384
18 Algebraic Methods
| =
L | |
(18.4)
of the many-electron the wave function. The eigenvector given by the matrix
L = i y2 represents the only triplet, and has nothing on the diagonal. Out
4
of the
= 6 states of the model, three are singlets. At half lling L is a
2
square m m matrix. A glance to (18.4) suggests that there would be much
to gain in complex problems if L were diagonal, because fewer numbers would
represent | , and one knows that L can be diagonalized if it is Hermitean.
If L can be taken Hermitean, it remains Hermitean when changing the basis.
Thus it is natural to ask if this advantage is free or we must pay for it.
Hermitean L
There is no magnetic eld, the Hamiltonian
is real and we take the basis
functions real as well. Then, if = , L, is an eigenstate of
H with eigenvalue , L also belongs to ; the transposed matrix LT has the
up and down spin amplitudes interchanged, and since this is a symmetry
operation it leaves the subspace of invariant. Hence, if L belongs to ,
this applies to L and for the Hermitean combinations L + L , i(L L ).
So we shall take L Hermitean1 . Hermiticity implies that we may impose the
normalization condition (14.60) | = 1 by writing T rL2 = 1. If all the
eigenvalues of L are non-negative, we write L 0. Then, we must have
T rL > 0 strictly, because otherwise L = 0 and the state vanishes. One can
infer about the spin of the many-body state2 if L 0. In fact, some diagonal
elements must be positive, and this holds true on the site basis like in (18.3);
therefore, is singlet, or has some mixture with a singlet. In the two-site
two-electron example one can see that for large negative U the eigenvalues of
L are of the same sign, while for large positive U they are of opposite signs:
attraction favors singlets.
In his proof, Lieb used the Scr
odinger Equation in the form (14.56):
ns Lns .
(18.6)
H : L K L + LKT + U
s
He had obtained (14.54)(for the real Hamiltonian case) from the variational
principle starting from the energy
1
;
(18.5)
L=
=
2
2
incidentally, this shows that the vanishing of the diagonal elements does not prevent
L from representing a singlet. However a singlet can give a non-zero contribution
to TrL while triplets and higher spins cannot.
2
It is not the sign of the wave function that brings information, of course, but
a non-vanishing TrL.
E = 2T r(KLL) + U
T r(Lnx Lnx ).
385
(18.7)
n.n.
dx dy
x,y
n.n.
= t
dy dx .
x,y
Having t for down spins and t for up spins is denitely ugly; so we change
signs to the orbitals in a sublattice (here the bipartite hypothesis enters)
replacing (18.8) by
dx = sign(x)cx .
(18.9)
Now, the transformed Hamiltonian reads
=t
H
n.n.
dx dy + cx cy U
dx dx nx + U N ; U > 0, (18.10)
x,y
0 t t0
= H(t, U ) + U = t U 0 t , U > 0
H
(18.11)
t 0 U t
0 t t0
are the same as those of H(t, U ).
With the above substitution of d for c, Sz of Equation (18.2) becomes
1
1
Sz =
([1 dx dx nx ]) = [ N N ]
2 x
2
and S + reads
S+ =
(x)dx cx .
(18.12)
(18.13)
Both are still conserved of course although they do not mean spin any more
in the transformed problem and are therefore called pseudo-spin.
386
18 Algebraic Methods
C= 2
and gets C LC = 2
, hence |L| = C 2
=
1 1
0 1
01
10
; this yields the ground state
C = 12
01
=
|x x | + |y y |
,
2
(18.14)
which is degenerate with (18.5) for U = 0 but is actually better for U < 0.
The question arises: is |L| really dierent from L? If it is, we get a new
ground state, namely the dierence, which we write in terms of the eigenvectors:
|i [|i | i ] i |.
(18.15)
R = |L| L =
i
Being a ground state, R > 0, unless R = 0. We now show that this is the case,
that is, |i | = i i. The latter equality must be true for some i, otherwise
L = |L| and we obtain the same ground state with a uninteresting (-)
sign in front; thus, the kernel of R is not empty, that is, RV = 0 for some
m-component vector V . The Schr
odinger equation (18.6) for R
K R + RKT + U
ns Rns = R.
(18.16)
s
387
Now since RV = 0 and Rnx V = 0 we nd RKV = 0, that is, every conguration connected to V by K is in the kernel; but any two conguration can
be reached from each other in a nite number of steps, that is K p has matrix
elements between them for some p. Finally the kernel has swallowed all the
Hilbert space! Thus, L = |L| (although, of course, we remain free to choose
L = |L|. Thus there is a singlet component in the ground state and L 0.
This implies that the ground state is unique. If L1 and L2 were dierent
ground states any combination L = L1 + L2 would be a ground state, but
we would be free to choose such that T rL = 0; this is impossible since any
ground state must be positive semidenite.
Repulsive Case
Now we assume H as in (18.1)with U > 0. The ground state is the transform
of the one with U < 0. (See Problem 18.1 for an example). This implies that
the eigenvalue is the same; the eigenvector is also the same, albeit it is written
in a reshued basis, and is therefore a unique singlet.
More generally, one can study bipartite lattices with sublattices A and B
with |A| and |B| sites per cell. Lieb showed that the ground state has spin
S = |B| |A| per cell; thus, the trivial Hubbard model has a singlet ground
state, while the three-band Hubbard model has S = 12 per cell.
d3 x
d3 p exp[
HH = J
N
Sn Sn+1 = H0 + HXY ,
(18.19)
n=1
H0 = J
N
n=1
z
Snz Sn+1
(18.20)
388
18 Algebraic Methods
while the so-called XY model hamiltonian, which has been studied by its own
sake, reads
N
"
J ! +
+
HXY =
Sn Sn+1 + Sn Sn+1
.
(18.21)
2 n=1
Let N r be the number of spins; HH is invariant for translations along
the chain (which implies conservation of a momentum K) and for rotation
in spin space (conservation of the total spin S and of its z component, Sz =
N N
, N = N N . For r = 0 the problem is trivial because
2
HH | . . . = E (N =0) | . . . , E (N =0) = J
N
4
(18.22)
and this is the only state. Here I assume ferromagnetic coupling, i.e. J > 0
for deniteness. Note that
H0 |1 , 2 , , N = {E (N =0) +
(J)
(2)N }|1 , 2 , , N
4
(18.23)
J
(|n + 1 + |n 1 ).
2
(18.24)
(N =0)
f (n + 1) + f (n 1)
f (n)
)f (n) = J
2
(18.25)
(18.26)
389
2j
, j integer
(18.27)
N
reecting the translation invariance of the problem. This represents a spin
wave excitation of a ferromagnet.
kj =
Two Magnons
For N = r = 2 it is natural write a basis {|n1 , n2 }, n2 > n1 specifying the
positions of spins. If n2 > n1 + 1, N = 4 and one nds:
(HH E (N =0) )|n1 ,n2
= 2|n1 , n2
J
|n1 +1,n2 +|n1 1,n2 +|n1 ,n2 +1+|n1 ,n2 1
;
+
2
(18.28)
let us take the scalar product f (n1 , n2 ) = E |n1 , n2 with the solution of
H|E = E|E , and write the recurrence formula:
(E E (N =0) )f (n1 , n2 )
= 2f (n1 , n2 )
J
f (n1 + 1, n2 ) + f (n1 1, n2 ) + f (n1 , n2 + 1) + f (n1 , n2 1)
(18.29)
.
+
2
This is satised by f (n1 , n2 ) = ei(k1 n1 +k2 n2 ) with energy eigenvalue
E = E (N =0) + J(1 cos(k1 ) + 1 cos(k2 ))
(18.30)
that would be the same as the sum of two magnon energies and would represent the superposition of two magnons if the momenta ki were given by
(18.27). However, the periodic boundary conditions do not apply separately
to the two momenta. Instead, we must have
or also
ei(k1 +k2 )N = 1,
(18.31)
eik1 N = ei , eik2 N = ei .
(18.32)
The interaction of the two magnons produces a relative phase shift, which
also changes the energy although the eigenvalue (18.30) is formally the sum
of two magnon energies. In terms of this (yet unknown) phase, we may write
N k1 = + 21 , N k2 = + 22 .
(18.33)
(18.34)
390
18 Algebraic Methods
the ordering 0 < n1 < n2 < N is needed otherwise the basis would be
over-complete.
The two contributions must enter with the same probability, A and B
dier by a phase (I have the right to use the same symbol, as I show
shortly) and we rewrite (18.34) as
f (n1 , n2 ) = ei(k1 n1 +k2 n2 + 2 ) +ei(k2 n1 +k1 n2 2 ) , 0 < n1 < n2 < N (18.35)
n2 < n1 + N.
(18.36)
This shift of the origin exchanges n1 and n2 , and the two terms in (18.35)
are therefore exchanged; the second term of (18.35) becomes the rst of
(18.36), but the phase /2 changes sign and an extra N k2 term appears in the exponent. These dierences must cancel because the boundary condition holds. Comparing with ( 18.35 ) we nd the conditions
n1 n2
n1 + 1 and the down spin in n1 cannot jump to the right and the one in n2
cannot jump to the left; Equation (18.29) must change because N = 2 and
f (n, n) does not occur in the equation since |n, n has no meaning, therefore
It is time worry about congurations like
(E E (N =0) )f (n1 , n1 + 1)
=
J
f (n1 1, n2 ) + f (n1 , n1 + 2)
= f (n1 , n1 + 1) +
.
2
(18.37)
1
[f (n1 , n1 ) + f (n1 + 1, n1 + 1)].
2
(18.39)
391
ei + 1
cot( ) = i i
2
e 1
(18.41)
k1
k2
2 cot( ) = cot( ) + cot( ).
2
2
2
(18.42)
The Bethe Ansatz equations (18.33 , 18.42), can be solved for k1 , k2 and
by combined analytic and numerical techniques for N some tens[72]. Most
solutions are scattering states with real momenta and a nite ; however when
one of the Bethe quantum numbers vanishes, the corresponding k and
also vanish, and there is no interaction. Other solutions with 1 and 2 close
to each other have complex momenta with k1 = k2 and represent magnon
bound states in which the two ipped spins propagate at a close distance.
Many Magnons
It is stunning that the Bethe ansatz works for any system size, and is generalized for any r. Again one chooses the origin and orders the spins with
ni < ni+1 ; a basis is {|n1 , n2 . . . nr }, specifying the positions of spins. We
rst seek a possible H eigenstate as a product of r magnon wave functions of
momenta kj . As the natural extension of (18.32) we allow each k to pick up
a phase from each of the others;
i
eiki N = e j=i ij , ij = ji .
(18.43)
Equation (18.33) becomes
N ki =
ij + 2i
(18.44)
j=i
and the integers i ranging from 0 to N-1 are Bethe quantum numbers. The
r! permutations P of the momenta yield degenerate r-magnon plane waves
that must be summed,so the Bethe Ansatz reads
!i r kPj nj + i PiPj "
2
j=1
i<j
(18.45)
e
f (n1 , . . . , nr ) =
P
392
18 Algebraic Methods
Again, we must consider two cases: 1) the reversed spins are all far away
2) there are neighboring reversed spins. In case 1) Schrodinger equation is
satised identically and yields the direct extension of (18.30) for the energy
eigenvalue:
r
(1 cos kj ).
(18.46)
E E0 = J
j=1
We can count our spins starting anywhere without changing the amplitude,
so we impose
f (n1 , , nr ) = f (n2 , , nr , n1 + N ).
This causes a permutation of the terms of the sum, and in each exponent the
term changes sign and an extra contribution proportional to N appears.
Exactly as in the two-magnon case, these changes are xed by (18.44) and
we may conclude that the phases are the same in (18.44) and (18.45).
Some changes occur in the Schr
odinger equation when there are neighboring reversed spins because N is reduced and some hoppings do not exist; the
same reasoning that took us to (18.40) can be extended directly and yields
eiij =
(18.47)
ij
ki
kj
= cot cot ,
2
2
2
i, j = 1, . . . , r.
(18.48)
The Bethe equations (18.44,18.48) are hard to solve for many ipped
spins in large systems. In recent years solutions have been obtained which
are far from trivial; among others, bound states of several magnons have
been reported.
The Bethe Ansatz is a rare example of exact solution of an interacting
many-body problem; it keeps the same form independent of the size of the
system. The root of the (relative) simplicity that allows this solution is onedimensionality: the evolution does not allow overtakings, and spins always
keep a xed order. However, without the ingenuity of Hans Bethe, the solution
could have remained undetected; who knows how many important model
problems are solvable, but still unsolved.
393
Sect. 9.8. The Hamiltonian, which can be shown [142] to be related to the
Kondo Hamiltonian, is:
N
2
H=
+ 2c
(xi xj ),
2
xi
i=1
i<j
c > 0.
(18.49)
(18.50)
It is:
(x1 . . . xi . . . xN ) = Q (x1 . . .
xi . . . xN ), 0 < xQ1 < . . . < xQN ,
. . . xi . . . xN ),
Q (x1 . . . xi . . . xN ) = P Q,P (x1
Q,P (x1 . . . xi . . . xN ) = [Q, P ]ei
kPi xQi
(18.51)
Here, P SN while k1 . . . kN denote dierent numbers, and [Q, P ] are amplitudes to be determined.
Thus each electron has 2 labels, the physical one is i, which is the order
along the line, while Qi depends on a ctitious (for indistinguishable particles)
and unobservable order, that is on the sector. However, if we remain in a
sector Q, we cannot describe any overtaking: when an electron i overtakes
the one on its right, the description automatically switches to a new sector.
Qi1
Qi Qi+1
Qi1
Qi+1 Qi
i1
i1
i+1
i i+1
i
Fig. 18.1. If a classical electron (triangle) overtakes another one they exchange
their physical labels and keep their sector labels. Since actual electrons are indistinguishable, one ends up in a neighbor sector Q = P (ii+1) Q, where P is a
transposition.
N
ki2
i=1
(18.52)
394
18 Algebraic Methods
Thus, Q =
P Q,P , where QP is [Q, P ] times a plane-wave state
where the ordering of electrons sector labels Q and the ordering of momenta
are xed. We are in a given sector Q as long as xQi increases as i increases,
but if two electrons with consecutive physical labels sit at the same x, we
are also in the sector Q = P (ii+1) Q, in which the images of i and i+1
are exchanged. This fact allows to impose sector boundary conditions: at the
boundary each plane wave must have the same amplitude in both neighboring
sectors. For example, in the Q sector the contribution to Q proportional to
[Q, P ] when xQ3 = xQ4 is
i(k +k )x +i
pPi xQi
i=3,4
Q,P = [Q, P ]e P3 P4 Q3
;
(18.53)
both in Q and in Q the contribution of P is summed to the contribution
of the permutation P = P (34) P bringing the same exponential:
i(k +k )x +i
pPi xQi
i=3,4
Q,P + Q,P = ([Q, P ] + [Q, P ])e P3 P4 Q3
. (18.54)
This amplitude must be the same as
Q ,P + Q ,P = ([Q , P ] + [Q , P ])e
i=3,4
pPi xQi
(18.55)
xQ4
dx(x) =
1
2
]xQ3 =xQ4 [
]xQ3 =xQ4 = c({xQi , xQ3 = xQ4 }).
xQ3
xQ3
(18.57)
395
Q ]xQ3 =xQ4 [
Q ]xQ3 =xQ4 = cQ ({xQi , xQQ3 = xQ4 }).
xQ3
xQ3
(18.58)
We need to write this in terms of the [P, Q] unknowns. In sector Q, xQ3 xQ4 ,
from (18.54)
i(k +k )x +i
pPi xQi
i=3,4
;
c({xi , x3 = x4 }) = c([Q, P ] + [Q, P ])e P3 P4 Q3
(18.59)
the contribution to [ xQ ]xQ3 =xQ4 proportional to [Q, P ] is
[
= ikP3 [Q, P ]e
Q,P
ikP3 xQ3 +i
i=3,4
kPi xQi
(18.60)
i=3,4
i=3,4
kPi xQi
pPi xQi
(18.61)
.
i=3,4
i=3,4
kPi xQi
kPi xQi
(18.63)
Q
P + Q P = i(kP4 [Q , P ] + kP3 [Q , P ])e
i=3,4
kPi xQi
.
(18.64)
(18.65)
(18.66)
396
18 Algebraic Methods
(x1 . . . xi . . . xN ) = (x1 . . . xi + L . . . xN ), i = 1, 2, N
(18.67)
which, like in the c = 0 case, determine the spectrum. However, unlike the
non-interacting problem, the N momenta are not independent. Before tackling this problem I show that the 1d Hubbard Model leads to similar equations.
18.3.2 The Hubbard Model in 1d
Consider N Fermions moving on a linear Hubbard chain of Ns sites
H =T
ci cj + U
ni ni ,
(18.68)
<ij>
f (i)y|i + 1 =
f (i)(i, y 1) = f (y 1);
(18.69)
(18.70)
Ns
x1 =1
Ns
...
...
xi =1
Ns
xN =1
f (x1 . . . xN )| x1 . . . xM xM+1 . . . xN ,
: ;< = :
;<
=
(18.71)
where xi is the site of electron i; more simply, we shall also write
|x1 . . . xM xM+1 . . . xN | x1 . . . xM xM+1 . . . xN .
: ;< = :
;<
=
N
[f (x1 , . . . , xi + 1, . . . xN ) + f (x1 , . . . , xi 1, . . . xN )]
i=1
i(1,M),j(M+1,N )
+U
(xi , xj )f (x1 , . . . , xi , . . . xj . . . xN )
(xi ,xj )
= Ef (x1 , . . . , xi , . . . xN ).
(18.72)
397
i(k2 x1 +k1 x2 )
fQ (x1 , x2 ) = [QP ]e
+ [QP ]e
fQ (x1 , x2 ) = [Q P ]ei(k1 x2 +k2 x1 ) + [Q P ]ei(k2 x2 +k1 x1 )
sector
Q : x1 x2 (18.75)
Q : x1 x2
If the spins are parallel, U does not act, and the two sectors are disjoint.
The Schr
odinger equation Equation (18.72) links the amplitude f (X) of
a given conguration X = (x1 , . . . , xi , . . . xj . . . xN ) to those that can be
reached from X in one step, that is, applying H once. If X RQ , applying
H we remain in RQ , if X is an interior point, that is |xi xj | > 1, i
= j;
otherwise X is a frontier point. We get dierent conditions from in the two
cases.
398
18 Algebraic Methods
X is an Interior Point
Let fQ (x1 . . . xN ) = P [QP ]ei n kPn xQn ; consider the contribution from a
particular P to the i = r term in the i sum of Equation (18.72): we obtain
the kinetic term
[fQ (x1 . . . xr + 1, . . . xN ) + fQ (x1 . . . xr 1, . . . xN )]
= (eikPr + eikPr )[QP ]ei n kPn xQn ;
(18.76)
N
cos(kj ).
(18.77)
j=1
Note that U enters indirectly, by modifying the k values. There are no interior points if the lling is too high, because then double occupations are
unavoidable; however (18.77) holds in general.
X is a Frontier Point
Consider any conguration in (18.74) with at least a double occupation on
a site that must be among the rst M and among the last M . For a given
Q, in (18.73) for some i, xQi = xQi+1 ; then, in Equation (18.74) for given
Q, P one nds at the exponent a contribution (kPi + kPi+1 )xQi . However,
if P is the permutation obtained from P by exchanging i and i + 1, the
resulting exponential is the same. For X Q, the permutations P and P
both contribute to fQ with the same exponential factor,
i
k
x
([Q, P ] + [Q, P ])ei(kPi +kPi+1 )xQi e m=i,i+1 Pm Qm .
The same exponential also arises from the permutation Q obtained from Q
by exchanging i, i + 1. Indeed, the contribution to fQ from P and P is
i
k
x
([Q , P ] + [Q , P ])ei(kPi +kPi+1 )xQi e m=i,i+1 Pm Qm .
We must ensure that f is single-valued, that is, {X RQ , X RQ } =
fQ (X) = fQ (X). This reproduces the condition (18.55)
[QP ] + [QP ] = [Q P ] + [Q P ].
(18.78)
399
by xQi xQi+1 . 3) (18.77) must hold in both sectors. Let site i be the
same as i + 1 (double occupation). In sector Q dened by xQi xQi+1 ,
the two electrons are those labelled Qi , Qi+1 ; there is superposition with the
sector Q = P (ii+1) Q where the permutations dier by Qi Qi+1 (P
is a transposition). The contribution to the Bethe function arising from a
permutation P of momenta is
i
k
x
[QP ]ei(kPi xQi +kPi+1 xQi+1 ) e m=i,i+1 Pm Qm ,
however in the case xQi = xQi+1 , also belonging to the sector there is mixing
of the permutations
P, P that dier by i i + 1; the exponential is the
same, e
m=,i,i+1
kPm xQm
+ [QP ]e
i kP xQi +kP
i
i+1
xQi+1
= [QP ]ei(kPi xQi +kPi+1 xQi+1 ) + [QP ]ei(kPi+1 xQi +kPi xQi+1 ) , (18.79)
such that the contribution to fQ arising from P and P is:
kPm xQm
PP
P i
m=,i,i+1
= P
e
, xQi xQi+1 .
Q
Q
(18.80)
e
, xQi xQi+1
(18.81)
Q
(same exponential as in (18.80) ) and
P
P
Q
= [Q P ]e
i+1
+ [Q P ]e
i kP xQ +kP
i
i+1
x Q
i+1
(18.82)
.
P
P
xQi xQi+1
Q
PP
Q xQi > xQi+1 .
(18.84)
We dene the set of the indices that in a given sector correspond to electrons that do not participate to double occupations: SQ := {i : xQi
=
400
18 Algebraic Methods
xQri1 , xQri+1 } Now, the rst member of (18.72) involves one sum on electrons; the contribution of those that do not participate to double occupations
is the correspondent piece of the energy eigenvalue (18.77):
N
PP
PP
QQ
(x1 . . . , xQ + 1, . . . xN ) + QQ (x1 , . . . , xQr 1, . . . , xN )
r
r,rSQ
= 2
N
PP
cos(kPr )QQ
.
(18.85)
r,rSQ
P
Since the eigenvalue is (18.77), the two-body function P
QQ must contribute,
and obeys in xQi = xQi+1 = x to (18.72) with eigenvalue 2 cos(kPi )
2 cos(kPi+1 ). Equation (18.72) grants that
P
PP
PP
[P
QQ (x + 1, x) + QQ (x 1, x) + QQ (x, x + 1)
P
PP
+P
QQ (x, x 1)] + U QQ (x, x)
P
= 2(cos(kPi ) + cos(kPi+1 ))P
QQ (x, x).
(18.86)
The 2 body problem is nested into the N body one in such the way that
everything generalizes. In view of (18.84), Equation (18.86) means
P
PP
[P
Q (x + 1, x) + Q (x 1, x)
P
PP
PP
+P
Q (x, x + 1) + Q (x, x 1)] + U Q (x, x)
P
= 2(cos(kPi ) + cos(kPi+1 ))P
Q (x, x).
(18.87)
ikPi+1
P
P
+ [Q P ]eikPi
Q (x + 1, x) [Q P ]e
ikPi
PP
Q (x 1, x) [QP ]e
+ [QP ]eikPi+1
P
P
[QP ]eikPi+1 + [QP ]eikPi
Q (x, x + 1)
PP
Q (x, x 1) [Q P ]eikPi + [Q P ]eikPi+1
P
P
[QP ] + [QP ]
Q (x, x)
(18.88)
we obtain
[QP ](eikPi + eikPi+1 + U ) + [QP ](eikPi + eikPi+1 + U )
= [Q P ](eikPi+1 + eikPi ) + [Q P ](eikPi + eikPi+1 ).
We eliminate [Q P ] using (18.78), and nd
(18.89)
[QP ] =
U
2i
[QP ]+(sin(kPi )sin(kPi+1 ))[Q P ]
U
sin(kPi )sin(kPi+1 )+ 2i
401
(18.90)
[1, 2, 3, . . . N , P ]
[Q2 , P ]
P =
(18.91)
...
[QN ! , P ]
Let 0 denote the column corresponding to the fundamental (or identity)
permutation I = {1, 2, 3, . . . N }. We cast the equations (18.66) in a more
convenient form,
[QP ] =
[Q P ] ij [QP ]
ic
, ij =
.
1 + ij
ki kj
(18.92)
(18.93)
To write down the column-switching operator Yija,b which performs a transposition of consecutive indices, we also need to introduce a row-switching
ba
operator PQQ
which acts on the Qi and exchanges a and b. Indeed,
Yija,b =
ba
ij + PQQ
aij + bij P ab , b = a + 1.
1 + ij
(18.94)
The upper indices of Yija,b give the position of the consecutive electron pair in
the list, the lower indices the identity of the elements; for example, for N=4,
3,4
exchanges the third and fourth elements in the list, provided that they
Y24
3
When it is clear from the context which are the i and j indices we are referring
to, we shall also continue to use the shorthand notation P for the permutation
of momenta obtained from P by exchanging ki and kj , and shall use P for the
column corresponding to it.
402
18 Algebraic Methods
3,4
2,3 3,4
are 2 and 4, giving Y24
3124 = 3142 ; moreover, Y14
Y24 3124 = 3412 , and
so on; any P can be obtained from 0 by applying products of Y operators.
The action of Yija,b is the same as the action of Pij on any permutation
P. From this very fact one that can prove some identities, that can of course
be checked starting from the denition of Y . Since Pij Pji = 1,
Yija,b Yjia,b = 1
(18.95)
(18.96)
Moreover, one can check that P12 P13 P23 = P23 P13 P12 ; in this way one nds
that
a,b b,c a,b
b,c
Yik Yij = Yijb,c Yika,b Yjk
.
(18.97)
Yjk
Here is the operator that exchanges both rows and columns:
ij
i,j
Xij = PQQ
=
Yij
ij
1 PQQ
ij
1 + ij
, j = i + 1.
(18.98)
This operator lets the electron i to overtake the electron j=i+1 (recall that
the phase factors have already been dealt with). For example if N = 3,
(18.99)
0 =
X1,3 0 =
,
X
X
=
2,1 1,3 0
kPi xQi
403
P ].
[Q,
(18.101)
Consider for instance the last but one component [{2, 3, 1}, {1, 2, 3}] of 0
in (18.99) and its evolution in (18.100): the boundary condition is
[{1, 2, 3}, {3, 1, 2}] = eik1 L [{2, 3, 1}, {1, 2, 3}].
(18.102)
(18.103)
ij
1 + PQQ
ij
1 + ij
(18.104)
(18.105)
but, since since we are looking for the spin part of the same wave function,
404
18 Algebraic Methods
j = eikj L .
(18.106)
These are N equations, but each yields all the eigenvalues since they can be
shown to commute; so we can limit ourselves to solve
X2N
XN
ZN = N , ZN = X1N
1,N .
(18.107)
(1)
= (2) ,
(18.108)
(3)
where the components can be interpreted as the amplitude that the overturned spin comes rst, second and third, respectively. In terms of the full
wave function for the three spins, (1) stands for [{, 2, 3}, {1, 2, 3}] = [{
, 3, 2}, {1, 2, 3}]; F (2) is the amplitude that spin 2 is down and is symmetric
in 1 and 3, and so on. The full 0 can be obtained from the reduced . Let
us work out4 the eigenvalue equation (18.107)
X23
.
Z3 = 3 , Z3 = X13
(18.109)
1
23
bij =
1 + 23
1 + 23
(18.110)
ij
we see that aij + bij = 1 and Xij
= aij + bij PQQ
; X23 (1) = (1), but
X23 mixes the other components:
4
Here and below this section involves some algebra which is quite elementary
but somewhat tedious; using a computer code is a good idea.
a23 b23
b23 a23
(2)
(3)
=
(3)
]
(2)[a23 + b23 (2)
405
,
.
(2)
(3)[a23 + b23 (3)
]
(18.111)
ic
2
ic
2
which means, the ratio of two consecutive components is obtained as the ratio
of two linear functions. Indeed, now we show that we can solve the boundary
condition eigenvalue problem by writing
1
k1 +ic/2
=
(18.112)
.
k2 ic/2
k1 +ic/2 k2 +ic/2
k2 ic/2 k3 ic/2
It holds:
a23 b23
b23 a23
(2)
(3)
=
3 (2)
(3)
2
,
,
(18.113)
kj + ic/2
;
kj ic/2
(18.114)
Thus,
1
Z3 = 3 (2) .
(18.115)
(18.116)
(18.117)
(3)
3 2 1
Now we can solve the eigenvalue equation by requiring that j = j and that
1 2 3 = 1;
(18.118)
this means that a translation by L of the whole system has no eect. Putting
together (18.106) and (18.114) one gets
eikj L =
kj + ic/2
kj ic/2
(18.119)
and with (18.118) one can determine and the momenta. The beauty of this
approach is that it can be fully extended.
406
18 Algebraic Methods
N Spins,1 Overturned
The above results extend immediately to arbitrary N ; thus,
(, y) =
y1
j=1
kj + ic/2
kj+1 ic/2
,
(18.120)
where the integer y species the position of the overturned spin along the
chain.
General Case
The above solution extends to any N and M via the following generalized
Bethe ansatz involving M dierent parameters:
=
a(P )(P1 , y1 )(P2 , y2 ) (PM , yM ),
(18.121)
P SM
with the same as before; here given two permutations P and P that dier
by the exchange of elements and + 1 the coecients obey[142]
a(P ) = a(P )
P P (+1) + ic
.
P P (+1) ic
(18.122)
kj + ic/2
.
kj ic/2
(18.123)
Problems
18.1. Work out the L matrix (Section 18.6) for the ground state of the repulsive H2 molecular model H(t, U ) (14.58) and for the canonically transformed,
attractive model (21.32) and compare the results for Ut = 7.
18.2. Work out the detailed solution for the Hubbard chain with N = 2
electrons. Verify the solution by writing down explicit energy eigenvalues for
Ns = 3.
Part VI
Appendices
Setting
b
L
= qx ,
c
L
d2 q =
= qy ,
U (s) = h
cL2
! a "2
a 3
d2 k
d2 k, (1.29) yields:
a
2
1 + k 2 e( s ) 1+k .
(19.1)
kx ,ky >0
yields
n
hc 2 L2 3
a
dz 1 + z e( s ) 1+z
3
4s
a
U (s) =
hc 2 L2
4s3
d
d
3
a
a
dz
e( s ) 1+z ;
1+z
en =
U (s) =
hc 2 L2
d
d
3
0
1
1e
(19.2)
1 = e11 ; thus
1
dz
.
1 + z e s 1+z 1
(19.3)
2
su
d
1
hc 2 L2
(19.4)
U (s) =
du !
"2 ;
2
d
s
esu 1
1
and putting
1
esu
1
du !
"2 =
u
s
e 1
Finally,
hc 2 L2
U (s) =
2
dx
es
d
d
1
(x 1)2
2
1
1
.
es 1
1
1
es 1
(19.5)
.
(19.6)
C3 = Z3 I
C3 C32
A1
1 1 1
1
C3v
A1
A2
E
I
1
1
2
2C3
1
1
1
C6v
A1
A2
B1
B2
E1
E2
I
1
1
2
2
(x, y)
C4v
A1
A2
B1
B2
E
I 2C5
2C52
1
1
1
1
1
1
2 2 cos 2 cos 2
2 2 cos 2 2 cos
I
1
1
1
1
2
2
C2
1
1
1
1
2
2
C2 2C
1
1
1
1
2 2 cos
2 2 cos 2
2C3
1
1
1
1
1
1
2C 6
1
1
1
1
1
1
I
1
1
1
1
C2v
A1
A2
B1
B2
2i
3
3v g = 6
1
z
1 Rz
0 (x, y)
C5v
A1
A2
E1
E2
C!v "
A1 ! + "
A2
E1 ()
E2 ()
=e
z
I
1
1
1
1
2
C2
1
1
1
1
2
2C4
1
1
1
1
0
C2
1
1
1
1
2 v
1
1
1
1
0
v
1
1
1
1
v
1
1
1
1
g=4
z
xy, Rz
x, Ry
y, Rx
2d g = 8
1
z
1
Rz
1 x2 y 2
1
xy
0
(x, y)
5v g = 10, = 2
5
1
z
1
Rz
0
(x,
y) "
!
0
xy, x2 y 2
3d
1
1
1
1
0
0
3v
g = 12
1
z
1
Rz
1
1
0 (x,
x , Ry
! y), (R
")
0
xy, x2 y 2
v
g=
1
z
1
Rz
0 ! (x, y) "
0
xy, x2 y 2
D2
A1
B1
B2
B3
I
1
1
1
1
C2z
1
1
1
1
C2y
1
1
1
1
C2x
1
1
1
1
g=4
x2 , y 2 , z 2
z, xy, Rz
y, xz, Ry
x, yz, Rx
412
D3
A1
A2
E
2C3 2C2
I
g=6
1 1
1
z 2 , x2 + y 2
1 1 1
z, Rz
2 1 0 (x, y) , (Rx , Ry )
E
1
1
1
1
2
2
1
1
1
1
2
2
2C6
1
1
1
1
1
1
1
1
1
1
1
1
C2
1
1
1
1
2
2C4
1
1
1
1
0
C2
1
1
1
1
0
2C2
g=8
1
z 2 , x2 + y 2
1
z
1
x2 y 2
1
xy
0 (x, y) , (Rx , Ry )
I
1
1
1
1
2
2
h
1
1
1
1
2
2
2C3
1
1
1
1
1
1
2S3
1
1
1
1
1
1
3C2
1
1
1
1
0
0
3v
g = 12
1
z2
1
Rz
1
1
! z
"
0 (x, y) , xy, x2 y 2
0
(xz, yz) , (Rx , Ry )
2C3
1
1
1
1
1
1
1
1
1
1
1
1
C2
1
1
1
1
2
2
1
1
1
1
2
2
3C2
1
1
1
1
0
0
1
1
1
1
0
0
3C2
1
1
1
1
0
0
1
1
1
1
0
0
i
1
1
1
1
2
2
1
1
1
1
2
2
2S3
1
1
1
1
1
1
1
1
1
1
1
1
D3h
A1
A2
A1
A2
E
E
D6h
A1g
A2g
B1g
B2g
E1g
E2g
A1u
A2u
B1u
B2u
E1u
E2u
I
1
1
1
1
2
2S6
1
1
1
1
1
1
1
1
1
1
1
1
h
1
1
1
1
2
2
1
1
1
1
2
2
3d
1
1
1
1
0
0
1
1
1
1
0
0
3v
g = 24
1
1
Rz
1
1
0 ! (Rx , Ry ) "
0
x2 y 2 , xy
1
1
z
1
1
0
(x, y)
0
Double Groups
C3v
A1
A2
E
E1/2
5
6
E E
C4v
A1
A2
B1
B2
E1
E2
E3
1
1
1
1
2
2
2
1
1
1
1
2
2
2
E E
1
1
2
2
1
1
1
1
2
2
1
1
C32
C3
2
3v 3v E
C3 E C3 E
1
1
1
1
1
1 1 1
1 1 0
0
1
1 0
0
1
1
i
i
1
1 i
i
C43 C2
2C2
2C2
C4
3
C4 E C4 E C2 E 2C2 E 2C2 E
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
0
2
0
0
0
2
0
0
0
2
2
0
0
0
As the components i,p and k vary, one nds a number of matrix elements
()
< i|Tp |k > that are all connected by the theorem.
Proof
R G, R = R
implies
i|Tq() |k = i|R RTq()R R|k =
()
i|R
Tq() Dqp
(R) R|k =
j|Tq() |r >
()
Dji (R) Dqp ()(R)Drk ()(R).
(21.2)
qjr
In order to use the GOT, we write the product of two D matrices as one
()
()
D. Actually,Dqp (R)Drk (R) is the direct product of two irreps,
()
()
()
(R)Drk (R) =
qr|s Dst t|pk ,
Dqp
st
st
(21.3)
414
()
()
Ng
js it ,
m
= j|Tq() |r
qjr
qr|s t|pk
st
NG
js it ,
m
that is
i|Tq() |k = T () i|pk ,
where the reduced matrix element
1
T () =
j|Tq() |r > qr|j
m qjr
does not depend on the components.
q.e.d.
(21.4)
Solutions
Problems of Chapter 2
2.1 np = Z1 T r(np ) has a denominator Z = k (1 + e(k ) ) and a
numerator k=p (1 + e(k ) )e(p ) , since p is lled in all states that
contribute.
2.2 One readily veries by computing the time derivative that
i t
U (t, t ) = U0 (t, t )
U (t, )H1 ( )U (, t )
h t
+
dt
(t t1 )T {[(t), A1 (t1 )] A2 (t2 )}
+(t t2 )T {A1 (t1 )[(t), A2 (t2 )] }
(21.5)
where the functions arise from (t ti ) factors (for (t) standing on the left
of Ai (ti ) and from (ti t) factors (for (t) standing on the right of Ai (ti );
hence one gets the commutators. The argument extends immediately to any
n yielding the term in and an extra term
n
(21.6)
q=1
2.4 One can thinkof the (t, t ) interval divided into N 1 intervals and
write eiH (t t) = n eiH (tn tn1 ) ; after dierentiating, one lets N .
Problems of Chapter 3
6
3.1 The ground conguration 1s 2s 2p has
= 20 states. The largest
3
ML is reached by (1+, 1, 0) thus ML = 2, MS = 12 . The resulting
2
416
Solutions
| ms1 ms2 |S = 1, MS = 0 =
2
2
2
and for |l1 = 2m1 l2 = 2m2 |L = 1, ML = 0 the following values:
m2 = 2 25
1
m2 = 1
10
m2 = 0 0
m2 = 1 m2 = 2
10
2
5
hence normalizing again (which is generally necessary when using ClebshGordan coecients)
|3 P ML = 0, MS = 0 = 110 {2|2+ , 2 | + 2|2 , 2+ | |1+ , 1 |
|1 , 1+ |}.
that L |0 = 2|1 , L |0 = 2| 1 .
So,
1
|3 P ML = 1, MS = 0 = L+ |3 P ML = 0, MS = 0 =
2
+
L+
1
1 + L2
1
|3 P ML = 1, MS = 1 = {2|2+, 1+ | 6|1+ , 0+ |},
10
Solutions
417
1
|3 P ML = 1, MS = 1 = {2|2 , 1 | 6|1 , 0 |},
10
1
|3 P ML = 0, MS = 0 = {2|2+ , 2 |+2|2 , 2+ ||1+ , 1 ||1 , 1+ |},
10
1
|3 P ML = 0, MS = 1 = {2|2 , 2 | |1 , 1 |},
5
1
|3 P ML = 1, MS = 1 = {2|1+, 2+ | 6|0+ , 1+ |},
10
1
|3 P ML = 1, MS = 1 = {2|1 , 2 | 6|0 , 1 |},
10
3.5 We found
1
|3 P ML = 0, MS = 1 = {2|2+ , 2+ | |1+ , 1+ |}.
5
Il is clearly a triplet since the spin conguration is , anf ML = m1 +m2 = 0.
In order to verify that it is a P state, we use
+
L2 = L21 + L22 + 2L1z L2z + L+
1 L2 + L1 L2
Problems of Chapter 4
4.1 Integrate
4.2 Let
d
d
! H
"
e [A, eH ] = eH [H, A] eH .
fmn (x, x , ) =
mn
418
Solutions
(z +
and is solved by
g(x , x, z) =
2m e
h2
2mz
h
2
(21.8)
|xx |
i z
(21.9)
|
= sign(x x ) ] and the band-edge singularity is similar to
[note that d|xx
dx
the discrete case (with one band edge, however).
4.4 In the continuous case, for 3d, the solution of (4.44), namely,
(z +
2 2
h
)g(r , r, z) = (r r )
2m
(21.10)
is readily veried to be
2mz
i
|rr |
h
2
e
m
g(r , r, z) =
.
|r r |
2
h2
Therefore,
1
Im limr r
(21.11)
z singularity.
1
4.5 U (r) = 12 n + S d2 r (g 1 (r, r , )+(E)
g (r, r , ))(r ), where
and E are the unperturbed and perturbed energy eigenvalues, respectively.
Problems of Chapter 5
5.3 We found already that
Gk0 () =
Vk0
G00 ().
k
(21.12)
Vk0
G0k (), k
= k ,
k
(21.13)
V0k Gkk
.
0 ()
(21.14)
1
k
|V0k |2
0 ()
(21.15)
Solutions
G0k () =
V0k
,
( k )( 0 ()) |V0k |2
419
(21.16)
and nally
Gkk () =
V0k
Vk0
.
k ( k )( 0 ()) |V0k |2
(21.17)
5.4 The Sz factor implies that 1) the scattering conduction spin direction is
the quantization axis for the impurity spin, 2) the singular scattering does not
ip the impurity spin 3) for S > 12 the scattering probability grows with the
square of the spin component mS 4) opposite mS gives opposite amplitude
(2) (2)
5) The resistivity goes with T rUif Uif .
Problems of Chapter 6
6.1 Averaging over a one-electron wave function,
e
j(x)
=
d3 x (x ) [px (x x ) + (x x )px ] (x ),
2m
but the rst term in the integrand can be rewritten (px (x )) (x
x )(x ) = i
hx (x ) (x x )(x ) and the expectation value is the current. Then
N
1
i)
d3 xA(xi ) j(x
H = 2mc
c
i
e
3
(21.18)
=
d x [A(x) px (x x ) + A(x)(x x )px ]
2mc
. The
where A(x) and px commute. Let us take the matrix element Hmn
rst term contributes
e
3
d x m (x ) d3 xpx A(x)(x x )n (x ).
2mc
Caution
is needed when integrating with the functions if there are operators.
3
d x(x x )(x ) = (x) holds even if contains operators, but must
stand on the left. So we cannot integrate over d3 x directly, but we can over
d3 x and we nd
e
d3 x m
(x )px A(x )n (x ).
2mc
The second contribution
e
3
d x m (x ) d3 xA(x)(x x )px n (x )
2mc
420
Solutions
Problems of Chapter 8
8.1
1) The characters of the representation (1) with one electron are
(E) = 5, (C2 ) = 1, (2C4 ) = 1, (2v ) = 3, (2d = 1. Applying the
LOT one nds (1) = 2A1 E B1 .
2) We choose a basis whose elements are determinants (i, j, k, m) |i
j k m | with, say, i > j, k > m, numbering the sites as in Figure 8.5.
2
5
= 100. One must determine how many
The number of congurations is
2
remain invariant or change sign under the operations of the Group. Under C2 ,
the only contributions come from (4, 2, 4, 2), (4, 2, 5, 3), (5, 3, 4, 2), (5, 3, 5, 3),
which remain invariant. Under x , one nds invariant congurations
(2, 1, 2, 1), (2, 1, 4, 1), (2, 1, 4, 2), (4, 1, 2, 1), (4, 1, 4, 1), (4, 1, 4, 2), (4, 2, 2, 1),
(4, 2, 4, 1), (4, 2, 4, 2), (5, 3, 5, 3) while the conguration that change sign are
(2, 1, 5, 3), (4, 1, 5, 3), (4, 2, 5, 3), (5, 3, 2, 1), (5, 3, 4, 1), (5, 3, 4, 2). The congurations (5, 2, 5, 2), (5, 2, 4, 3), (4, 3, 5, 2), (4, 3, 4, 3) remain invariant under
one of the d reections. Proceeding in this way, one nds the characters of
the reducible representation (4) for 4 particles:
C4v I C2 2C4
A1
1 1 1
A2
1 1 1
B1
1 1 1
B2
1 1 1
2 2 0
E
(4) 100 4 0
2 v
1
1
1
1
0
4
2d g = 8
1
z
1
Rz
1 x2 y 2
1
xy
0
(x, y)
4
Applying the LOT one nds (4) = 15A1 11A2 24E 13B1 13B2 .
8.2 The characters of (4) are 1296, 16, 0, 64, 16; hence, (4) = 184A1
144A2 320E 176B1 152B2 .
Problems of Chapter 9
9.1
In E E = A1 A2 E all the irreps are contained. Using the set of
matrices (7.44),
E
10
01
+
C3
1
3
2
2
3
1
2
2
,+
C32
12
23
3
2
12
c
,
a + b , +
,
1
3
1
3
1 0
2
2
2
2
3
3
0 1
1
1
2
Solutions
421
y+x 3
3
, C32 x = x+y
, C32 y =
2
2
x+y 3
y+x 3
x+y 3
, 3 x = 2 , 3 y
x, 1 y = y, 2 x = 2 , 2 y =
2
the projection operator P (A1 ) = RG R, one nds
Using C3 x = x+y2
, C3 y =
3
, 1 x =
y+x 3
and
2
y+x2
=
x2 +y2
1 3
PA1 x1 x2 = x1 x2 +( x1 +y21 3
)( x2 +y22 3 ) + ( x1 +y
2 )(
2
x1 +y1 3
x2 +y2 3
x1 y1 3 x2 y2 3
+x1 x2 + (
)(
)+(
)(
)
2
2
2
2
= 3 (x1 x2 + y1 y2 ) .
.
(21.19)
A1 =
2
9.2
|x1 y2 y1 x2
,
(21.20)
|1 A2 =
2
to be multiplied for the singlet spin function. Therefore the non-vanishing
coecients are: 1 A1 |Ex1 Ey1 = 1 A1 |Ey1 Ey1 = 12 .
9.3 Using the set of matrices (7.44),
E
10
01
+
C3
23
2
3
21
2
,+
C32
12
23
3
2
12
c
,
a + b , +
,
1
3
1
3
1 0
2
2
2
2
3
3
0 1
1
1
2
x1 +y1 3 x2 +y2 3
)(
)
P(E,y) x1 x2 = x1 x2 12 (
2
2
1 x1 +y1 3 x2 +y2 3
2(
)( 2
)
2
1 x1 +y1 3 x2 +y2 3
+x1 x2 2 (
)(
) 12 ( x1 y21 3 )( x2 y22 3 )
2
2
= 32 (x1 x2 y1 y2 ) ;
P(E,x) x1 x2 = 0. Hence,
|1 Y =
|x1 x2 |y1 y2
.
2
Thus,
@1 1
@ 1
1
X 1 E1 x1 E2 x2 = 1 X 1 E1 y1 E2 y2 = .
2
422
Solutions
(J) () =
sin[(J + 12 )]
sin( 2 )
(21.21)
which which generalizes (9.20) above and results from using the identity
J
m=J
1
x
cos(mx) = csc( ) sin[(J + )x].
2
2
One nds
(3/2) () =
sin(2)
= 4 cos()) cos( ).
sin( 2 )
2
(21.22)
The reections and all the improper rotations can be written like products
iR . Proceeding in this way,
3
C43 C4 C2 2C 2 2C2
D 4 E E
2C2 E
C4 E C4 E C2 E 2C 2 E
A1 1 1
1
1
1
1
1
1
1
1
1
1
A 2 1 1
B 1 1 1 1 1
1
1
1
1
1
1
B 2 1 1 1 1
E 1 2 2 0
0
2
0
0
E 2 2 2 2 2 0
0
0
E 3 2 2 2
2
0
0
0
3/2
4 4 0
0
0
0
0
2
0
0
0
5/2 6 6 2
Problems of Chapter 10
10.1 This result is most easily obtained
in the time representation, keeping
Solutions
Vkq Fqk (, , t, t )( ; t, t ),
i
423
(21.23)
Problems of Chapter 11
11.1
R
(d)
(i)U (ilkj)(i)U (kjil)
0
iklj
dt1
t1
dt2
(21.24)
1
.
1 G0
(21.25)
i t
H0 Vef f (1))(1, 2) and we may rewrite (11.81) as G(0)1 = G1 + ,
that is,
G1 (1, 2) = (i
(21.26)
Problems of Chapter 12
12.1 One gets
2
E
n
(3 2 ) 3
2
23
3 2 (
t1
t2
t3
(iU )4 t
dt
dt
dt
dt4 (e2iV (t1 +t2 t3 t4 ) )2
1
2
3
28
0
0
0
0
1 + 2e2iV t
1 U 4 5 e4iV t 4e2iV t
( )
iV t
=
. (21.27)
32 V
512
128
424
Solutions
Problems of Chapter 13
13.1 Symbolically, one writes Equation (13.135)
J = (L) f (L) (g r g a ) + (L) g <
but for the other electrode one writes
J = (R) f (R) (g r g a ) (R) g < ;
the solution arrives eliminating g < algebraically; this is legitimate for every
innitesimal energy interval d.
Problems of Chapter 14
14.1 Dening the combinations
p = um + us , p = um us , r = um + ius , s = um ius ,
one obtains
Gmn (E) =
prob1
14.2 One gets the moments and nds an = 0 and
b2n = W 2
n2
.
4n2 1
(21.28)
W2
.
4
(21.29)
+
A
A
A
b1
a12 b2 + a13 b3
a12 a23 b3 + a13 a32 b2
=
+
(21.30)
.
D1
D12
D123
We may proceed by noting that D123 = D12 D312 , D312 = a33 D12 = D1 D21 ,
D21 = a22 a23 a32 /D312 , and we have all the ingredients to apply 14.86 and
carry on the calculation.
Solutions
425
Problems of Chapter 16
16.1 Using (16.15),(16.16), the exponent transforms to
hk 2
mv 2
mvx
)x + t(
+
]
h
2m
2h
but this must be written in terms of x . Thus one obtains
i[(k
(x , t) = eik x ih
(k )2 t
2m
with hk = h
kmv. Thus the plane-wave becomes a plane-wave with Galilean
transformation of the momentum.
Problems of Chapter 17
17.1 Table III shows how the irreps of G split in C4v .
G
A1
A1
B2
2
B
1
2
1
2
3
4
1
2
3
4
1
2
3
4
1
2
C4v
A1
A1
B2
B2
2B2
2A1
A1 + 2B1
A1 + 2B1
2A2 + B2
2A2 + B2
A1 + B1 + E
A1 + B1 + E
A2 + B2 + E
A2 + B2 + E
A2 + B1 + 2E
A2 + B1 + 2E
A1 + B2 + 2E
A1 + B2 + 2E
2A1 + 2B1 + 2E
2A2 + 2B2 + 2E
Table III. Reduction of the irreps of Optimal Group G of the 4 4 model in the
point Group.
426
Solutions
Problems of Chapter 18
18.1 In the repulsive case, the Hamiltonian (14.58) reads
U t t 0
t 0 0 t
H=
t 0 0 t.
0 t tU
(21.31)
1
with Q =
, q = (U + 16 + U 2 ). For U = 7, =
2
2
16+U
+U 16+U
0.181492 0.683418
0.531129, L =
.
0.683418 0.181492
In the attractive case (21.32)
0 t t0
= H(t, U ) + U = t U 0 t , U > 0
H
(21.32)
t 0 U t
0 t t0
on the basis ( a , b , a , b )) where = b and = a for the
new Fermions. The transformed matrix is
=Q
2 q
L
q 2
1
=
with Q
, q = (U + 16 + U 2 ). For U = 7, =
2
2
16+U
U 16+U
0.683418 0.181492
=
0.531129, L
.
0.181492 0.683418
18.2 With 2 fermions, Q = P = {1, 2}, Q = P = {2, 1}, let k1 < k2 denote
the momenta, and we may write
fQ (x1 , x2 ) = [QP ]ei(k1 x1 +k2 x2 ) + [QP ]ei(k2 x1 +k1 x2 ) x1 x2
fQ (x1 , x2 ) = [Q P ]ei(k1 x2 +k2 x1 ) + [Q P ]ei(k2 x2 +k1 x1 ) x1 x2
con
[QP ] =
U
[QP ] + {sin(k1 ) sin(k2 )} [Q P ]
2i
sin(k1 ) sin(k2 ) +
U
2i
(21.33)
and so on. For parallel spins U cannot act, sectors Q and Q are disjoint and
the coecients are all determined by [QP ] = [Q P ] = [Q P ] = [QP ] and
normalization.
Solutions
427
For antiparallel spins, the periodicity conditions are also needed. In sector
Q, with 1 x1 x2 Ns let the amplitude be fQ (x1 , x2 ). Let the electron
1 sit at be site x1 ; if we decide to relabel x1 + Ns that site, the amplitude
does not change, but we are now in sector Q . So,
fQ (x1 , x2 ) = fQ (x1 + Ns , x2 ).
(21.34)
Now,
[QP ]ei(k1 x1 +k2 x2 ) + [QP ]ei(k2 x1 +k1 x2 )
= [Q P ]ei(k1 x2 +k2 (x1 +Ns )) + [Q P ]ei(k2 x2 +k1 (x1 +Ns ))
implies
(21.35)
(21.36)
(21.37)
(21.38)
[Q P ] = [QP ].
(21.39)
Integers 1 2 are Bethe quantum numbers that label the state with
[QP ] = [QP ]ei .
(21.40)
= [Q P ]e
(21.41)
428
Solutions
This yields
(21.42)
(21.43)
sin(k2 ) sin(k1 )
sin(k2 ) sin(k1 ) +
U
2i
U
2i
(21.44)
i sin(k)
i sin(k) +
U
4
U
4
ei + 1
.
cot
= i i
2
e 1
(21.45)
(21.46)
This leads to
U cot( ) = 4 sin(k).
2
For odd n we may also write
4
tan( n ) = sin(k).
2
2
U
(21.47)
(21.48)
Introducing
0 (p) = 2 arctan(
2p
),
U
(21.49)
we get
= n + 0 (2sin(k)).
One can verify the above results for Ns= 3, in the singlet sector.Direct di2
2
(2 times), 1+U+ 92U+U
agonalization yields the eigenvalues 1+U 92U+U
2
2
2
2
(2 times), 2+U+ 36+4U+U
and for the ground state 2+U 36+4U+U
.
2
2
Solutions
429
These results are reproduced by the Bethe ansatz with Equation (18.77).
From = 0, that is k = 3 the solution of (21.47)
= 6 arccos[
U
5
+
8 16
36 + 4U + U 2
]
16
leads to the ground state and to the last but one. The other states arise from
= 1 and = 2, and each choice yields both eigenvalues.
References
432
31.
32.
33.
34.
35.
36.
37.
38.
39.
40.
41.
42.
43.
44.
45.
46.
47.
48.
49.
50.
51.
52.
53.
54.
55.
56.
57.
58.
59.
60.
61.
62.
63.
64.
65.
66.
References
E.N. Lassettre, Suppl. Radiation Research 1, 530 (1959)
U. Fano,Phys. Rev.124 1866 (1961)
P.W. Anderson,Phys. Rev.124 41 (1961)
J. Kondo, Prog. Theor. Phys. 32, 37 (1964).
J. Kondo, Theory of dilute Magnetic Alloys, in Solid State Physics, Ed. Seits
23, (1969).
C. Verdozzi, Y. Luo, and Nicholas Kioussis, Phys. Rev.B 70, 132404 (2004)
E. H. Lieb, Phys. Rev. Lett. 62, 1201 (1989).
B. I. Lundqvist, Phys. Kondensierten Materie 9, 236 (1969).
David C. Langreth, Phys. Rev. B 1, 471 (1970).
R. Brout and P. Carruthers, Lectures on the Many-Electron Problem (WileInterscience, NewYork, 1963)
Gerald D. Mahan, Phys. rev. B 11 4814 (1975 )
Gerald D. Mahan, Many-Particle Physics, Plenum Press, New york and London
(1990)
H. Ehrenreich and M. H. Cohen, Phys. Rev.115, 786 (1959)
E. N. Economou : Greens functions in Quantum Physics, Springer - Veriag,
1979. Second edition 1983
Michele Cini and Andrea Dandrea, J. Phys. C 21, 193 (1988)
M.Cini, A. DAndrea and C. Verdozzi, International Journal of Modern Physics
B9, 1185 (1995)
M.Cini, Il Nuovo Cimento 9D, 1515 (1987)
A. Blandin, A. Nourtier and D.W. Hone, J. Phys. (paris) 37, 369 (1976)
Adolfo Avella, Ferdinando Mancini and Roland M
unzner, Phys. Rev. B63,
645117 (2001)
Ferdinando Mancini and Adolfo Avella, cond-mat/0006377
Ferdinando Mancini and Adolfo Avella,The Hubbard model within the equation
of motion approach, advances in Physics 53, 537 (2004)
J. Lindhard, Kgl. Danske Videnskab. Selskab, Mat.-Fys. Medd.,28, no.8 (1954)
Lars Hedin, Phys.Rev.139, A796 (1965).
G.Onida,L.Reining and A.Rubio, Rev. Modern Phys. 74, 601 (2002)
J.M.Luttinger and J.C. Ward,Phys. Rev. 118, 1417 (1960.
G.Baym and L. Kadano, Phys. Rev. 124, 287 (1961);
G. Baym, Phys. Rev. 127, 1391 (1962).
C. Lanczos, J. Res. Natl. Bur. Stand. 45, 255 (1950).
R. Haydock, V. Heine and M.J. Kelly, J. Phys. C 5, 2845 (1972), R. Haydock, Solid State Physics Ed. Eherenreich, F. Seitz and D. Turnbull (London,
Academic Press, 1980), R. Haydock and C.M.M. Nex, J. Phys. C 18, 2235
(1985).
F. Cyrot-Lackmann, J. Phys. C:Solid State Phys. 5, 300 (1972); F. CyrotLackmann, M.C. Desjonqueres and J.P. Gaspard, J. Phys. C:Solid State Phys.
7, 925 (1974)
R. Landauer, IBM J. Res. Dev. 1, 233 (1957); Philosof. Mag. B21, 863 (1970)
C. Caroli,R. Combescot, P. Nozieres and D.Saint-James, J.Phys. C:Solid State
Phys. 4, 916 (1971); ibidem 5, 21 (1972)
Michele Cini, Phys. Rev. B22, 5887 (1980)
Yigal Meir and Ned S. Wingreen, Phys. Rev. Letters 68, 2512 (1992).
M.V. Berry, Proc. R. Soc. Lond. A392, 45 (1984)
C. Verdozzi and M. Cini, Phys. Rev. 51, 7412 (1995).
References
433
434
References
103. G.Stefanucci and C.O. Ambladh, Phys. Rev.B 69, 195318 (2004)
104. Michele Cini, Phys. Rev. B43,4792 (1991); Michele Cini,R. Del Sole, Jang Guo
Ping and L. Reining, Calculation of the linear and nonlinear optical properties
of Si(111)1x1-As, Proceedings of the Epioptic meeting, Berlin, June 1991.
105. Y.R. Shen, The principles of Nonlinear Optics, John Wiley and Sons, New
York (1984).
106. Schuda, C R Stroud Jr and M Hercher, J. Phys. B: Atom. Molec. Phys. 7,
L198 (1974)
107. P.Auger, J. Physique Radium 6, 205 (1925)
108. J.J.Lander, Phys. Rev. 91, 1382 (1953)
109. Claudio Verdozzi, Michele Cini and Andrea Marini, Jnl. Electron Spectroscopy
and Related Phen VOL.117-118, 41-55 (2001)
110. C.J.Powell,Phys. Rev. Letters 30, 1179 (1973)
111. L.I. Yin, T.Tsang and I. Adler,Physics Letters 57A,193 (1976); Phys. Rev. B
15, 2974 (1977)
112. G.A. Sawatzky, Phys. Rev. Letters 39, 504 (1977)
113. Knotek and Feibelman, Phys. Rev. 1978
114. J.E. Ingleseld, J. Phys. C:Solid State Phys. 14, 3795 (1981)
115. S. Crampin, J. Phys.: Condens. Matter 16, 8875 (2004)
116. Richard D. Mattuck, A Guide to Feynman Diagrams in the Many-Body Problem,Dover, New York
117. John Inkson, Many-Body theory of Solids - An Introduction,Plenum Press New
York 1984
118. R. P. Feynman,Statistical Mechanics - A set of lectures, W.A. BENJAMIN,
INC, Reading, Massachussetts (1972)
119. G. Vignale and Mark Rasolt, Phys. Rev. Letters 59, 2360 (1987)
120. Robert van Leeuwen, Nils Erik Dahlen, Gianluca Stefanucci, Carl-Olof Almbladh and Ulf von Barth, Lectures Notes in Physics, Springer Verlag, 2005 (to
appear); also cond-mat/0506130.
121. P. Hohenberg and W.Kohn, Phys. Rev. 136, B864 (1964)
122. W. Kohn and L.J. Sham, Phys. Rev. 140, A1133 (1965)
123. L.J. Sham and M. Schl
uter, Phys. Rev. Letters 51, 1888 (1983)
124. Michele Cini and A. DAndrea, J. Phys. C 16, 4469 (1983)
125. R.M. Dreizler and E.K.U. Gross, Density Functional Theory, Springer Verlag
1990
126. Ulf von Barth, Niels Erik Dahlen, Robert van Leeuwen, and Gianluca Stefanucci,Phys. Rev. B72, 235109-1 (2005)
127. Andrea Marini and Michele Cini, Phys. Rev. B 60, 11391 (1999)
128. M. Cini and V. Drchal, J. Phys.: Condens. Matter 6, 8549 (1994);M. Cini and
V. Drchal, J. Electron Spectrosc. Relat. Phen.72, 151 (1995)
129. M.J. Stott and E. Zaremba, Phys Rev A 21, 12 (1980)
130. A. Zangwill and P. Soven, Phys Rev A 21, 1561 (1980)
131. E. Runge and E.K.U. Gross, Phys. Rev. Letters 52, 997 (1984).
132. J.R. Schrieer and D. C. Mattis, Phys. Rev 140, A1412 (1965);B. Kj
ollerstr
om, D.J. Scalapino and J.R. Schrieer,Phys. Rev 18, 665 (1966)
133. V. Galitzkii, Soviet Phys. JETP 7, 104 (1958).
134. M. Cini and C. Verdozzi, Solid State Commun.57, 657 (1986);M. Cini and C.
Verdozzi, Nuovo Cimento 9D, 1 (1987).
135. A. Pernaselci and M. Cini, J. Electron Spectrosc. and Relat. Phen. 82,79
(1996).
References
136.
137.
138.
139.
140.
141.
142.
143.
144.
145.
146.
147.
148.
149.
150.
151.
152.
153.
154.
155.
156.
157.
158.
159.
160.
161.
162.
163.
164.
165.
166.
167.
168.
169.
170.
171.
172.
435
436
References
173. Adalberto Balzarotti, Michele Cini, Enrico Perfetto and Gianluca Stefanucci,
J. of Phys.: Condens. Matter 16 R1387 (2004)
174. W.Kohn and J.M. Luttinger, Phys. Rev. Letters 15, 524 (1965)
175. J.G. Bednorz and K.A. M
uller, Z. Phys. B64, 189 (1986)
176. RW Richardson, Phys. Rev. 141, 949 (1966)
177. G. Fano, F. Ortolani and A. Parola, Phys. Rev. B 42, 6877 (1990); A. Parola,
S. Sorella, M. Parrinello and E. Tosatti, Phys. Rev. B 43, 6190 (1991); G.
Fano, F. Ortolani and A. Parola, Phys. Rev. B 46, 1048 (1992).
178. C. A. Balseiro, A. G. Rojo, E. R. Gagliano and B. Alascio, Phys. Rev. B 38,
9315 (1988).
179. J. E. Hirsch, S. Tang, E. Loh and D. J. Scalapino, Phys. Rev. Lett. 60, 1668
(1988).
180. M. Ogata and H. Shiba, J. Phys. Soc. Jpn. 57, 3074 (1988).
181. J. Bardeen, L.N. Cooper and J. R. Schrieer, Phys. Rev. 108, 1175 (1957):
J.R. Schrieer, The theory of superconductivity, Perseus Books, Reading, Massachussets (1964)
182. E. Cappelluti, B. cerruti and L. Pietronero, Phys. Rev.B 69 , 161101 (2004)
183. M. S. Hybertsen, M. Schl
uter and N. Christensen, Phys. Rev. B 39, 9028
(1988); M. Schl
uter, in Superconductivity and Applications, edites by H.
S. Kwok et al. (Plenum Press, New York, N. Y.) p. 1 (1990).
184. F. C. Zhang and T. M. Rice, Phys. Rev. B 37, 3759 (1988)
185. M. Cini, A. Balzarotti, R. Brunetti, M. Gimelli and G. Stefanucci, Int. J. Mod.
Phys. B 14, 2994, (2000).
186. M. Cini and A. Balzarotti, Solid State Comm. 101, 671 (1997); M. Cini, A.
Balzarotti, J. Tinka Gammel and A. R. Bishop, Il Nuovo Cimento D 19, 1329
(1997).
187. M. Cini, A. Balzarotti and G. Stefanucci Eur. Phys. J. B 14, 269 (2000).
188. E. Perfetto, G. Stefanucci, and M. Cini, Phys. Rev. B 66, 165434 (2002); E.
Perfetto , G. Stefanucci , M. Cini Eur. Phys. J.B 30, 139 (2001).
189. M. Cini, E. Perfetto and G. Stefanucci, Eur. Phys. J. B 20, 91 (2001).
190. Michele Cini, Adalberto Balzarotti, Raaella Brunetti, Maria Gimelli and Gianluca Stefanucci, International J. Modern Phys. 14, Nos. 25-27, pag. 2994
(2000)
191. C.P. Lund, S.M. Thurgate and A.B. Wedding,Phys Rev. B55, 5455 (1997)
192. Stefano Bellucci, Michele Cini, Pasquale Onorato and Enrico Perfetto, J.
Phys.: Condens. Matter 18,S2069 (2006)
193. G.W. Fernando, A.N. Kocharian, K. Palandage, Tun Wang and J.W. Davenport, Cond-mat/0608579
194. David J. Griths, Introduction to Quantum Mechanics,Prentice Hall , upper
Sadle River, New Jersey 07458 (1994)
195. A.Callegari, M. Cini, E. Perfetto and G.Stefanucci, E Eur. Phys. J. B 34, 455
(2003)
196. M. Cini, G. Stefanucci and A. Balzarotti, Eur. Phys. J. B 10, 293 (1999).
197. M. Kociak et al., Phys. Rev. Lett. 86, 2416 (2001). A. Kasumov et al., Phys.
Rev. B 68, 214521 (2003)
198. F. Mancini and A. Avella, Advances in Physics 53, 537 (2004)
199. Satoro Odashima, Adolfo Avella and Ferdinando Mancini, Phys. Rev. B 72,
205121 (2005)
200. G. Bressi, G. Carugno, R. Onofrio and G. Ruoso, Phys. Rev. Lett. 88 041804
(2002)
References
437
Index
C3v , 137
C3v , 137
GL(n), 133
O(n) , 134
O+ (3), 201
SL(n) , 134
SO(3), 145
SO(4), 145
SU (5), 201
U (1), 145
SCF method, 33
++ , 81
1-2-3 theory, 270
Abelian Group, 133
Abrikosov-Suhl peak , 96
accidental degeneracy, 371
adiabatic electronic Hamiltonian, 188
adiabatic ionization energy, 115
adiabatic theorem, 359
advanced Greens functions, 57, 288
AES, 109
ammonia, 158
Anderson, 285
anti-Stokes, 335, 340
APECS, 109, 267, 270
APEX, 121
ARUPS, 109
atomic levels, 42
atomic terms, 41
Auger, 210
Auger eect, 48
Auger Electron Spectroscopy, 109
Auger matrix elements, 52
Auger Photoelectron Coincidence
Spectroscopy, 109
Auger selection rules, 50
Auger spectroscopy, 56
440
Index
Index
Kanamori solution, 122, 127
Keldysh, 210, 286, 294
Klein-Gordon equation , 65
Knotek-Feibelman mechanism, 121,
122, 210, 212, 270
Kohn-Sham equations, 281, 282
Kondo, 95, 393
Kondo Model, 99
Kondo peak , 96
Kondo temperature, 105
Koopmans, 33
Kramers, 340
Kramers Theorem, 172
Kubo, 75, 285
Kubo formulae, 73
L-S multiplets, 40
Lagrange theorem, 139
Lanczos, 313, 325
Landauer, 304
Langreth, 288
laser, 345
LCAO, 93
Lehmann, 76, 77
lesser and greater Greens functions, 57
Lie Group, 133, 144
Lieb, 324, 383
Ligand group orbitals, 161
Linked Cluster Theorem, 225, 226, 231
linked diagrams, 225
Lippmann-Schwinger equation, 70
little Group, 176
Local Density Approximation, 281
local density of states, 66, 86
LOT, 152
Low Density Approximation (LDA),
264
magnon, 388, 391
many-body Hamiltonian, 12
master equation, 305
Matsubara, 60, 290
mean eld, 31
Migdal theorem, 255
mixed coupling scheme , 52
mixed valence , 95
molecular orbitals, 158
moments, 322
441
442
Index
star of k , 176
Stokes, 335, 340
SU(n), 134
subgroup, 134
superconducting ux quantization, 368,
369, 378
Symmetric Group, 199
symmetry, 135
symmorphic Group, 174
t matrix, 70
temperature Greens function, 59
temperature Greens function , 62
tensor, 340
Index
Wicks theorem, 219
Wigner Eckart theorem, 45
Wigner matrices, 145
Zener model, 98
443