Esplora E-book
Categorie
Esplora Audiolibri
Categorie
Esplora Riviste
Categorie
Esplora Documenti
Categorie
UNIVERSITA
Dottorato di Ricerca in Scienze Chimiche
Sturmian Orbitals in
Quantum Chemistry
Andrea Caligiana
Coordinatore:
Relatore:
Ottobre 2003
Ringraziamenti
Desidero ringraziare sentitamente numerose persone,
che hanno contribuito allo svolgimento del mio dottorato
ed alla realizzazione di questa tesi: innanzitutto il Prof.
Aquilanti, la cui disponibilita a venirmi incontro e sostenermi nei momenti dicili e stata ancora piu importante
di qualunque competenza scientica un relatore possa
trasmettere. Lo stesso dicasi per Simona (Prof.ssa Cavalli), che non ha mai smesso di infondermi ducia e
di aiutarmi a conoscere me stesso. Una parte considerevole dei risultati presentati e dovuta alla preziosa
collaborazione con il Prof. John Avery dell'Universita
di Copenhagen. Il prof. Tarantelli mi ha tante volte
trasmesso parte delle sue conoscenze non solo al Dipartimento di Chimica, ma anche durante lunghi viaggi a
bordo del bus n.9. L'estate del 2002, trascorsa al Dipartimento di Chimica di Cambridge(UK), e stata un periodo stimolante e piacevole, oltre che scienticamente
produttivo: ringrazio il Prof. Handy che mi ha permesso
di lavorare nel suo gruppo di ricerca e di alloggiare al
St. Catharine's College, ed i suoi collaboratori Rudolph
Burcl e David Tew.
Dopo i `senior', passo a menzionare coloro con cui
(oltre al luogo di lavoro) ho condiviso svariate serate
divertenti,... Andrea Beddoni, Andrea Lombardi, An-
Contents
Introduction
5
1 Hydrogenic Elliptic Orbitals, Coulomb
10
16
24
31
39
48
53
60
62
67
69
75
76
80
81
2
2.4.2 Zeeman Basis Sets . . . . . . . . . . . . . .
2.5 The elliptic bases of O(4) . . . . . . . . . . . . . . .
2.5.1 Spheroelliptic Coordinates and Basis sets . .
2.5.2 Elliptic Cylindrical Coordinates of Type I and
II . . . . . . . . . . . . . . . . . . . . . . . .
2.6 Summary and concluding remarks . . . . . . . . . .
Bibliography . . . . . . . . . . . . . . . . . . . . . . . .
82
87
88
90
95
99
Introduction . . . . . . . . . . . . . . . . . . . . . .
Free atoms . . . . . . . . . . . . . . . . . . . . . . .
Atoms in external elds . . . . . . . . . . . . . . .
Electronic integrals between Sturmian congurations
3.4.1 Overlap Integrals . . . . . . . . . . . . . . .
3.4.2 Matrix elements of one-electron operators .
3.4.3 Matrix elements of two-electron operators .
3.5 On the choice of the `energy' in the denition of the
basis set . . . . . . . . . . . . . . . . . . . . . . . .
3.6 Examples . . . . . . . . . . . . . . . . . . . . . . .
Bibliography . . . . . . . . . . . . . . . . . . . . . . . .
106
107
111
112
113
114
116
120
122
124
148
151
151
156
163
3
Appendix . . . . . . . . . . . . . . . . . . . . . . . . . . 164
Bibliography . . . . . . . . . . . . . . . . . . . . . . . . 166
Conclusions
170
171
174
181
187
190
192
197
Bibliography . . . . . . . . . . . . . . . . . . . . . . . . 200
Introduction
The fundamental scope of quantum chemistry is the determination of the wavefunctions for the stationary states of electrons
in the presence of xed nuclei, eventually including the eect of
external elds and perturbations. Then, according to the BornOppenheimer approximation, the electronic wavefunctions may be
used for the construction of potential energy surfaces, on which
the motion of the nuclei can be studied, by quantum or classical
techniques.
The electronic wavefunctions are obtained solving a time independent Schrodinger equation by methods that include the use
of sets of monoelectronic functions (orbitals) as basis sets for the
approximation of the solutions. The kind of functions used as electronic orbitals aects strongly the convergence rate and accuracy
of numerical procedures.
The only chemical system whose Schrodinger equation is analytically solvable is the hydrogen atom. In the early days of quantum
chemistry, scientists tried to use hydrogenic functions (i.e. the solutions of the Schrodinger equation for the hydrogen atom) and
5
6
similar functions as orbitals to approximate more complex wavefunctions, incurring in the inconvenient of complicated two-electron
integrals. In the last decades, instead, linear combinations of Gaussian functions have been used as electronic orbitals, replacing all
the other types of basis sets: Gaussians have allowed a faster calculation of two electron integrals, and consequently an extension
of quantum chemistry towards larger systems, permitted also by a
great improvement of computer performances. At the present time,
the main goals of researchers are the implementation of calculations
for larger and larger systems, and also a better understanding of
excited states.
This Doctoral Thesis is devoted to the study of a type of functions, the Sturmian hydrogenic orbitals, which so far have received
very little attention in the eld of quantum chemistry. Hydrogenic
Sturmians are formally similar to the well-known hydrogenic orbitals, but their use has important and peculiar implications that
are worthwhile investigating. We have considered the applications
of Sturmians in ab initio quantum chemistry, without entering the
eld of density functional theory. The Thesis is structured in three
parts.
7
function in conguration space; consequently, there are different types of hydrogenic Sturmians in conguration space.
Symmetry properties and orthonormal transformations among
them are studied.
Chapter 2 deals with the properties of Sturmians in momentum space. In the reciprocal space, Sturmians are strictly
related to hyperspherical harmonics. This property is very
interesting by itself and also because it suggests the application of angular momentum techniques in quantum chemistry.
There are dierent types of hydrogenic Sturmians in momentum space; as in the previous chapter, Symmetry properties
and orthonormal transformations among them are studied.
The relationship between the two reciprocal spaces is also
discussed.
The second part of this thesis (Chapters 3,4 and 5) is dedicated to the applications of Sturmian basis sets in quantum
chemistry.
In chapter 3 the application of Sturmian orbitals as a basis set for atomic structure calculations is investigated, also
including the eect of external elds.
In chapter 4 there is a treatment of Sturmian formalism applied to one electron molecules according to the MO-LCAO
method.
Chapter 5 deals with an original method for exploiting Stur-
8
mians in a valence bond approach to the study of molecules.
The passage to momentum space is used in order to solve
some multi-center one-electron integrals
Chapter 1
Hydrogenic Elliptic Orbitals, Coulomb
Sturmian Sets and Recoupling
Coecients among Alternative Bases
Summary
The non-relativistic Schrodinger equation for the Coulomb problem is separable in four dierent coordinate systems in conguration
space: alternative sets of orbitals for the hydrogen-like atoms correspond to each of them and permit to obtain Sturmian sets, useful as
complete orthonormal expansion bases in atomic and molecular calculations. In this chapter the fundamental properties of the already
known hydrogenic orbitals (the familiar polar, the parabolic and the
rarely treated spheroidal sets) are resumed; then we discuss some
properties of the spheroelliptic orbitals, which have been so far practically ignored. We pay particular attention to the symmetries of
the dierent orbital sets and to the relationships between them, and
9
10
we order them in a complete scheme which exhibits passages from
one to the other through explicitly derived orthogonal transformations. Within this context we insert the study on the conservation
of parity in the passage from the polar set to the parabolic one. We
also show that - except for the spheroidal set - all alternatives for
hydrogenic wave functions induce irreducible representations of the
point group D2h , the `quaternion group'. This has relevance for the
discussion of connections between these sets and the corresponding
ones in momentum space, presented in Chapter 2.
1.1 Introduction
The Schrodinger equation for the hydrogen atom was solved in
the early days of quantum mechanics by using polar coordinates
(r; #; '). The other coordinate systems suitable for separating and
solving the equation are less known; each of them leads to hydrogenic orbitals with dierent properties. A systematic search on this
topic was completed in the 70's exploiting the relations between
separation of variables for dierential equations and Lie group theory [1,2]. In spite of the apparent simplicity of the hydrogen atom
as a physical system , its quantum-mechanical treatment turns out
to be wide and complicated; the reason can be found in its exceptional symmetry, which is the symmetry of the (hyper)-sphere
S 3 , embedded in a four-dimensional Euclidean space. Our interest
towards this study is due not only to the scientic quest for under-
11
standing properties of the only atom of the periodic table which
can be treated exactly by an analytical approach, but also to the
important aspect that the various types of orbitals are candidates
as basis sets for the numerical solution of other atomic and molecular systems, and are therefore basic building blocks for quantum
chemistry.
The physical quantum states for a hydrogenic atom with nuclear
charge Z are described by the wave functions which are solutions
of the Schrodinger equation:
(H^
E)
1 2
r
2
Z
r
=0
(1.1)
where r is the distance nucleus-electron. This is an exactly solvable two-body problem, at variance with all the other atoms of the
periodic table. The hydrogenic hamiltonian operator H^ has a discrete spectrum (bound states, negative E ) and a continuous one
(unbound states, positive E ), and one has to take both to get a
complete set. In view of diculties which arise when working with
continuous basis sets, it is better to associate a discrete set to the
solution of the quantum problem in Eq. (1.1). Such a complete set
of orbitals is called Coulomb Sturmian, the nomenclature originating from the fact that Eq. (1.1) is a Sturm-Liouville problem.
The quest for completeness lead many authors to introduce sets
which are closely related to usual orbitals, but for which completeness holds with no need of continuum states, an early example being
the "Natural Spin Orbitals" by Shull and Lowdin [3]. General Stur-
12
mian functions were introduced by Rotenberg [4, 5], starting from
the Schrodinger equation for the variable r in a radial-eld system
in the three dimensional physical space. The most important example is the Coulomb problem, for which Sturmian functions are
dened as solutions of the equation
"
1 d 2d
r
2r2 dr dr
l(l + 1)
+ E0
2r2
Zn
R (r) = 0
r n;l
(1.2)
p0 = ( 2E0 )1=2
(1.3)
13
and u1;0;0(r; #; ') is the wave function for the orbital 1s; no other
un;l;m(r; #; ') corresponds to a physical wave function, but a complete set can be explicitly dened. Coulomb Sturmians obey a
peculiar orthonormality relation [6]:
Z
1
p
un;l;m(r; #; ')dr = 0 nn0 ll0 mm0
r
n
(1.4)
n;l;m
r0 ) (# #0 ) (' '0 )
(1.5)
In the next section we will further discuss the apparent coincidence
and the dierences between Coulomb Sturmian eigenfunction sets
and the usual hydrogenic orbitals.
The hamiltonian in Eq. (1.1) can be easily formulated, by using
a Cartesian coordinate system centered on the nucleus, but it is
known that these coordinates lead to a non-separable dierential
equation. The problem of the separation of variables is ubiquitous
in quantum mechanics, because in the study of systems with many
degrees of freedom one tries to reduce the original multidimensional
problem in the hope of getting problems of lower dimensionality,
which are simple enough to be solved analytically or numerically,
14
and allowing us to better understand the physical process and to
refer to simpler cases. Separation of variables can be exact or only
approximate. Exact separation can be viewed as symmetry separation, since it is associated to geometrical and dynamical symmetries
of the system [10, 11]; this is the type of variable separation which
applies to the study of the hydrogen atom as a two-body problem
and thus will be under focus in this chapter. Symmetry separation
usually involves the introduction of separation constants to isolate
each variable, and for each of them one gets a dierential equation where only one degree of freedom appears. In other words, a
partial dierential equation becomes a set of (uncoupled) ordinary
dierential equations.
For most quantum systems, which are made of three or more
bodies, an approximate separation can be performed, based on the
assumption that the motion associated to a group of variables can
be treated as the other ones were frozen; the "fast" variables produce an eective potential for the motion of the "slow" variables,
which will have to be treated accordingly. The semiclassical nature
of this adiabatic separation procedure has been elucidated [12, 13].
The Born-Oppenheimer separation is an example, as well as the
hyperspherical approach to the few-body problem.
In the next sections, we will consider the four coordinate systems
which allow the exact separation of the Schrodinger equation. For
each coordinate system the properties of the corresponding Sturmian basis vary and the ranges of problems for which they are
15
appropriate basis sets dier accordingly.
Sturmian sets in conguration space have hyperspherical harmonics basis sets as their counterparts in momentum space. Care
will be taken to develop the treatment to prepare for the following
chapter, where hyperspherical symmetry is exploited in the momentum space framework. The correspondence is best understood
making reference to symmetry, and thus to group theory. The second order commuting operators which appear as constants of the
motion in the treatment of hydrogen atom in momentum space - our
next chapter will deal with the latter - are invariant with respect to
the inversion operation: as a consequence, momentum space Sturmians induce irreducible representations of the point group D2h [2]
(Schon
ies notation is used throughout this thesis, see Ref. [14]).
Thus, only conguration space Sturmians inducing irreducible representation of D2h have a counterpart in the momentum space corresponding to the same set of quantum numbers. For this reason
we pay special attention in the present chapter to the behaviour
of the conguration space Sturmians in the point group D2h . This
correspondence will be treated in the following chapter, where previous problems such as the Coulson's "intractability of spheroidal
orbitals" [15] will be shown to arise from the dierent behaviour of
commuting operators resulting from separation of variables in the
two reciprocal spaces with respect to inversion of parity.
Notations and properties of polar and parabolic Sturmians, having been recently discussed (and also extended to the general d-
16
dimensional hydrogen atom case) [16], are brie
y resumed in Sec. 2,
while in Sec. 3 we implement the requirement of parity conservation
in the orthogonal matrix which connects the two sets. Spheroidal
Sturmians are described in Sec. 4 (the corresponding hydrogenic
orbitals have been lucidly illustrated by Sung and Herschbach [17]).
Spheroelliptic Sturmians and their properties are introduced in Sec.
5, where their expansion coecients in terms of the other sets are
given as solutions of a three-term recursion relationship, which also
allows us to obtain the `missing label', i.e. the additional quantum
number characterizing the set (this is typical of elliptic coordinates
and basis sets, an ample variety of which will be encountered in
the next chapter). This work leads to the complete list of Coulomb
Sturmian sets and of the recoupling coecients among alternative
bases: the corresponding connection scheme will be exhibited in
the concluding Section.
17
coordinates (see g. 1a) are the most natural choice for solving eq.
(1.1). Writing the polar Sturmian as a product of a radial factor
Rn;l (r) and an angular factor Yl;m (#; ') enables us to separate the
coordinate r from the others. After introducing the separation constant l(l + 1), we obtain two dierential equations: the rst one is
eq. (1.2), and the second one
"
1 @
@
1 @2
sin #
+ 2
Y (#; ') = l(l + 1)Yl;m(#; ')
sin # @#
@#
sin # @'2 l;m
(1.6)
coincides with the equation for the free motion of a particle on
the surface of the three-dimensional sphere S 2 ; it is known that the
wave functions of that system are the spherical harmonics Yl;m (#; ') =
Nl;m Pljmj(cos #) eim' (Nl;m is the normalization factor [18]). The
explicit expression of the polar Coulomb Sturmians in conguration
space is:
"
#
l 1)! 1=2
2n[(n + l)!]3
(n
un;l;m(r; #; ') = (2p0 )3=2
(1.7)
l 2l+1
0 r) Ln l 1 (2p0 r) Yl;m(#; ')
p r (2p
0
where p0 =
2E0 , n = 1; 2; 3; : : :, l = 0; 1; : : : ; n 1 and l
m l; L2nl+1l 1 (2p0 r) is a Laguerre polynomial [19]. The use of p0
instead of the more traditional energy E0 serves to emphasize the
connection with the treatment of the same problem in momentum
space, which will be presented in the following chapter. It is important to stress that Eq. (1.7) also represents hydrogenic orbitals:
the dierence between orbitals and Sturmians is that in the former
18
^l2
np0
un;l;m(r) = 0
r
i
l(l + 1) un;l;m(r) = 0
(^lz m) un;l;m(r) = 0
(1.8)
(1.9)
(1.10)
In Eq. 1.8 the product np0 gives the "charge" Zn . To these three
relationships we add
P^ un;l;m(r) = un;l;m(r)
(1.11)
where P^ is the operator that performs the inversion of all the electronic coordinates with respect to the nucleus, and = ( 1)l .
The following real linear combination of polar orbitals are conveniently dened ( = 1):
(1.12)
(1.13)
(1.14)
19
Table 1 - Number of polar (Eq. 1.12) or spheroelliptic functions for
given l, and labels.
; !
l=1
2
3
4
5
6
...
even l
odd l
+1,+1
-1,+1
1
2
2
3
3
4
...
l=2 + 1
(l + 1)=2 (l
+1,-1
-1,-1
0
1
1
1
1
1
1
2
2
2
2
2
2
3
3
3
3
3
...
...
...
l=2
l=2
l=2
1)=2 (l + 1)=2 (l + 1)=2
(1.15)
(1.16)
Therefore the functions un;l;jmj are seen to induce irreducible repre^ C^2 (x); C^2 (y ); C^2(z )g
sentations in the point group D2h = D2 Ci (D2 = fE;
^ P^ g). This is also known as the "quaternion group".
and Ci = fE;
For symmetry operations and groups, we use the notation of Ref.
[14]. Table 1 shows the number of un;l;jmj functions for given l;
and values.
The separation of variables in the Schrodinger equation written
in polar coordinates is always possible for the motion in any radial
eld. In the case of Coulomb eld, separation of variables is also
possible in parabolic coordinates (Fig. 1a).
These coordinates have two points of reference: one is on the
20
z
z
) /2
2
1+
P ()
( 1 2 ) /2
P ()
(+) A
(a)
( 1 2)
1/2
A (+)
(b)
x
=cost
Figure 1.1:
The part (a) of this gure shows polar and parabolic coordinates. The denition of the polar coordinates with respect to the cartesians
is x = r sin # cos '; y = r sin # sin '; z = r cos #. The relatioships of parabolic
coordinates with the cartesians and the polars are x = (1 2 )1=2 cos '; y =
(1 2 )1=2 sin '; z = 12 (1 2 a nd 1 = r(1 + cos #); 2 = r(1 cos #); ' = '.
The part (b) serves for the denition of spheroidal coordinates: = a+ b where
1 1 and = a b where 1 1. ' is the angle between the plane
ABP and a xed plane through AB; 0 ' < 2 . Therefore if the y axis is s
een as pointing downward in the plane of the drawing, ' is as in the part (a).
21
nucleus, and the other is at innite distance along a direction that
in Fig. 1a coincide with the z axis. The explicit expression for the
parabolic sturmians is, according to the notation in [20]:
1
2
3
0
p0p12 e
Lm(n+
1+
2
(1.17)
m
im'
m 1) (p0 1 ) L (n m 1) (p0 2 ) e
1
2
np0
Un;;m (r) = 0
r
K^ z Un;;m (r) = 0
^lz m Un;;m (r) = 0
(1.18)
(1.19)
(1.20)
K^ =
1 ^
l p^
2
p^ ^l
Zr
=p0
r
(1.21)
22
The vectors K and l are orthogonal in classical mechanics, and for
^ = 0. Since ^l is invariant
the quantum mechanical operators ^l K
with respect to inversion, while p^ and r change sign, eq. (1.21)
^ also changes sign under inversion; thus K^z doesn't
implies that K
commute with P^ and the parabolic functions [Eq. (1.17)], being
eigenfunctions of K^ z , do not have a well dened parity. We will
come back to this point in the following section, both because this
is crucial in establishing relationships with the momentum space
treatment and also because for applications it is often useful to
have a basis of denite parity.
The trasformation matrix between the polar set jn; l; mi [Eq.
(1.7)] and the parabolic set jn; ; mi [Eq. (1.17)] has to conserve the
quantum numbers n and m, then it is made of blocks of dimension
n jmj because for xed n and m, there exist n jmj polar orbitals
and n jmj parabolic orbitals. Within a phase factor, the matrix
elements - which coincide with the overlaps of the functions of the
two sets - are particular Clebsch-Gordan (CG) coecients [18, 22,
24, 25] in which the two coupling vectors have the same modulus
[26]:
jn; l; mi
Pn jmj
= (n jmj 1)
( ) (n +m 1) cn;m
l; jn; ; mi
1
2
(1.22)
1); 2 + 2 2
2 + 2 j l; mi
This relationship shows that the connection between the polar and
parabolic basis can be seen as a change of coupling schemes [17],
^ ).
involving the angular momentum operators k^ = 12 (^l K
cn;m
l;
= h 21 (n
m ; 1 (n
1);
23
An alternative view of these matrix elements stems from the
Regge symmetry [25, 27] for the Clebsch-Gordan coecients:
h 12 (n 1); 12 (m ); 12 (n 1); 21 (m + )jl; mi =
h 12 (n + m 1); 2 ; 12 (n m 1); 2 jl; 0i
(1.23)
The last form (a Clebsch-Gordan coecient where the projection
of the orbital angular momentum is zero) suggests that can be
interpreted as a helicity quantum number [21]. Indeed the choice
of the reference frame for which the component of the orbital angular momentum vector ^l is zero amounts to change the quantization
axis on the plane of the orbit, where the Runge-Lenz vector lies.
This choice is reminiscent of analogous ones which occur in different contexts: we recall the transformation between space-xed
and body-xed frames in molecular collisions [28], the transformation between the Hund's cases (e) ! (c) in molecular spectroscopy
and atomic scattering [29, 30] and the passage between symmetric
and asymmetric coordinates in the hyperspherical treatment of the
three-body problem [31, 32].
In general, the CG coecients enjoy two basic three-term recurrence relationships with respect to angular momentum and to
the projection quantum numbers [33, 34]. For the CG coecients
in Eq. 1.22, we have with respect to l:
v
u
u
t
v
u
u [(n
[n2 l2 ][l2 m2 ] n;m
1)2 l2 ][l2 m2 ] n;m
t
cl 1; cn;m
+
cl+1; = 0
l;
(2l 1)(2l + 1)
(2l + 3)(2l + 1)
(1.24)
24
and with respect to :
1q 2
[n
4
(1.25)
Eqs. 1.24 and 1.25 can be seen from a matricial point of view:
the CG coecients that appear there are eigenvectors of tridiagonal
matrices. Eq. 1.24 corresponds to the matrix representation of K^ z
by the set jn; l; mi, the eigenvalues being ; eq. 1.25 corresponds to
the matrix representation of ^l2 by the set jn; ; mi, the eigenvalues
being l(l + 1). Here we see a manifestation of properties which can
be viewed from the relationships between CG coecients and Hahn
polynomials of a discrete variable [35]. For these polynomials and
their duals, there is an exchange in the role of the discrete variable
and their degree, and Equations like (1.24) and (1.25) can be also
seen as the nite dierence equations respectively either in l or .
25
a matrix is centrosymmetric (a matrix A of dimensions d d is
centrosymmetric if Aj;k = Ad+1 j;d+1 k ), implying the existence of
a simple similitude transformation which factorizes it in two blocks
[36]. Accordingly, the set jn; ; mi transforms into a new set of
eigenfunctions of the parity operator P^ , which commutes with ^l2 .
Block-diagonalization occurs because matrix elements of ^l2 between
functions which have opposite parity vanish. The explicit form of
the new basis of denite parity, which will turn out to be connected
by a simple relationship with the originary basis, can be obtained
by projecting out the even and odd components of the parabolic
Sturmians in Eq. (1.17).
As we can see from Figure 1, inverting the position vector r
of the electron amounts, in parabolic coordinates, to exchanging
1 and 2 and turning ' into ' + . By summing or subtracting
Un;;m (1 ; 2 ; ') and Un;;m (2 ; 1 ; ' + ), we get the even (U + ) and
the odd (U ) components of the latter ( = 1, as before):
1
U (1 ; 2 ; ') = p [Un;;m (1 ; 2 ; ') Un;;m (2 ; 1 ; ' + )] =
2
(1.26)
1
= p [Un;;m (1 ; 2 ; ') ( )m Un;;m (2 ; 1 ; ')] =
2
q
m
h
1
= p Nn;;m p0 1 2 e p
Lm(n+ m 1) (p0 1 )
2
i
Lm(n m 1) (p0 2 ) ( )m Lm(n m 1) (p0 1 )Lm(n+ m 1) (p0 2 ) eim'
0
1
2
1
2
1+
2
1
2
1
2
26
parity parabolic orbitals are linear combinations of two orbitals
which have opposite values of :
1
Un; jj;m = jn; ; m; i = p [Un;;m ( )m Un;
2
;m ]
(1.27)
(1.28)
(1.29)
jn; l; mi =
n X
jmj 1
=0 or 1
( ) (n +m 1) Cl;n;m;
jj jn; ; m; i
1
2
(1.30)
where the step of the sum over is two units; it starts from 0 or 1,
according to whether n jmj 1 is even or odd: indeed, in the ket
jn; ; m; i of dened parity, can only be positive or zero. The
desired orthonormal transformation matrix has to be factorized in
two blocks distinguished by the sign of the parity , at xed values
of n and m. We have
1 + ( )l n;m
q
Cl;n;m;
=
cl;jj
jj
2(1 + ;0 )
(1.31)
27
For each parity, columns are now labeled by jj and , instead of
as in Eq. (1.22), and the quantities Cl;n;m;
jj are combinations of CG
coecients in Eq. (1.22) which dier only for the sign of :
Cl;n;m;
jj =
h
i
1
m cn;m
cn;m
+
(
)
l;
l;
2(1 + ;0 )
(1.32)
0
B
B
B
B
@
u4;0;0
u4;1;0
u4;2;0
u4;3;0
C
C
C
C
A
B
B
B
B
B
@
3
2
1
2
1
2
q
1
2
q
1
5
1
5
1
2
q
1 1
2 5
1
2
q
3 1
2 5
1
q2
1 1
2 5
1
2q
3 1
2 5
1
q2
3 1
2 5
1
q2
1 1
2 5
10
CB
CB
CB
CB
C@
A
U4; 3;0
U4; 1;0
U4;1;0
U4;3;0
1
C
C
C
C
A
(1.33)
28
Block-diagonalizing according to the conservation of parity, one has:
0
B
B
B
B
@
u4;1;0
u4;3;0
u4;0;0
u4;2;0
1
C
C
C
C
A
B
B
B
B
B
B
@
3q 101
0
0
1
10
1
q 10
3 101
0
0
0
0
p1
2
p1
2
0
0
p1
2
p1
2
0
C
CB
CB
CB
CB
C@
A
U4;3;0
U4;1;0
U4+;1;0
U4+;3;0
1
C
C
C
C
A
(1.34)
Eq. (1.31) denes analogues of the CG coecients which conserve the parity; we will refer to them as parity-conserving ClebschGordan coecients (pCG). In the following we will show how to
work out the three-term recurrence relationships enjoyed by pCG
coecients. Indeed, the eigenvectors of the matrix of K^ z2 in the
basis jn; l; mi are eigenfunctions also of H^ and ^lz : in fact since
[K^ z2 ; H^ ] = [K^ z2 ; ^lz ] = 0, the matrix elements of functions which differ in n and/or m are zero, and n and m are preserved as good
quantum numbers. Therefore, the matrix is diagonalized by the
coecients of the orthogonal transformation between jn; l; mi and
jn; ; m; i: those coecients are exactly pCG-coecients. We will
see that the matrix is tridiagonal, and therefore the pCG coecients
enjoy a three-term recurrence relationship.
As we have seen, [K^ z2 ; P^ ] = 0, and the parity of jn; l; mi is ( 1)l ,
so we expect that only matrix elements hn; l0 ; mjK^ z2 jn; l; mi with l
and l0 having the same parity can be dierent from zero. This
implies that the matrix is factorized in two blocks, for even and
29
odd l. In the l-th row only three elements are dierent from zero:
m ][n (l+1) ]
[l m ][n l ]
hn; l; m j K^ z2 j n; l; mi = [(l+1)(2l+1)(2
+
l+3)
(2l 1)(2l+1)
2
[n2
m]
2
30
1
1 2
(n 1) + (m2 2 ) l(l + 1) Cl;n;m;
jj +
2
2
(m + + 1)2 ] [n2 (m 1)2 ] Cl;n;m;
j+2j = 0 (1.37)
1q 2
[n
4
is identical to the recurrence for the CG (Eq. 1.25), apart from the
range of .
Finally, we complete this section by building-up explicitly real
parabolic orbitals which enjoy the symmetry properties of the group
D2h : they are
Un;;jj;jmj = jn; ; m; ; i = (2i 1 ) 1=2 Un; jj;m Un; jj;
(1.38)
(1.39)
(1.40)
(1.41)
(1.42)
For xed n and jmj, there exist n jmj functions of type Un;;jj;jmj.
Table 2 indicates how many of these functions there are for each
couple of and labels at a given n jmj value. From the fact
that eq. (1.22) is invariant with respect to the sign of m, it follows that the coecients of the transformation between jn; l; m; i
and jn; ; m; ; i coincide with the pCG coecients, i.e. with the
coecients of the transformation between jn; l; mi and jn; ; m; i.
31
Table 2 - Number of D2h parabolic functions (Eq. 1.38) for given
and labels and n jmj xed.
; !
even m
odd m
+1,+1
-1,+1
(n
jmj)=2 (n jmj)=2
+1,-1
(n
0
jmj)=2 (n
-1,-1
0
jmj)=2
32
monograph on spheroidal functions.
In order to solve the Coulomb problem in spheroidal coordinates one writes the spheroidal Sturmians T (; ; ') as a product
( ) N ( )('), and introduces two separation constants, m2 e f .
Dening
p
Z
E0 1=2
= 0 = n
2
2
2n
we one can write the separated equations as:
d2
2
2 = m (1.43)
d'
"
#
"
#
d
d
m2
2
2
2
(1 )
+ f + 2n
= 0 (1.44)
d
d
1 2
The equation in is formally identical to the one in , but the
interval of denition of the variable is dierent (see Fig. 1a). Eq.
(1.43) also occurred in the separation in the polar variables # and
' (see Eq. 1.6); its normalized solutions are again (') = p12 eim' ,
and the T (; ; ') are eigenfunctions of ^lz .
Let us consider Eq. (1.44). Since for ! 1 the asymptotic
solution is e , we demand for the unknown function ( ) the
form:
nX1
( ) = e
cl Pljmj( )
(1.45)
l=jmj
33
relationship for the cl is obtained [17, 22]:
(n + l + 1)(l + m + 1)
(n l)(m l) f + 2 l(l + 1)
+cl
cl+1
=0
(2l 1)
2
(2l + 3)
(1.46)
This amounts to say that, after xing n and jmj, it is necessary to
diagonalize a tridiagonal matrix of order n jmj to get the n jmj
spheroidal orbitals; the eigenvalues are the separation constants
f associated to each orbital. As a consequence, at variance with
what happens for the polar and parabolic basis sets, neither the
spheroidal eigenfunctions nor the eigenvalues f can be written as
a "closed" expression. Symmetrizing the recurrence relationship of
the cl (see Appendix) one has [22]:
cl 1
v
u
u
t
h
[n2 l2 ][l2 m2 ]
gl 1 + l(l + 1) f
(2l 1)(2l + 1)
2 gl +
v
u
u
t
(^l2 + 2K^ z ) = (f + 2 )
by expanding
(1.48)
nX1
l=jmj
gl jn; l; mi
34
besides obviously those of H^ and ^lz . The classical observable which
corresponds to the former operator is a constant of the motion in
the case of a two-centre Coulomb potential [45].
We list the eigenvalue equations which completely characterize
the spheroidal basis:
!
1 2 p20 np0
r + 2 r Tn;f;m (r) = 0
n = 1; 2; 3 : : : (1.49)
2
h
i
(^l2 + 2K^ z ) (f + 2 ) Tn;f;m (r) = 0
(1.50)
^lz m Tn;f;m (r) = 0
jmj = 0; 1; : : : ; n 1(1.51)
So, in order to obtain spheroidal orbitals without solving Eq. (1.44)
we can calculate the coecients of the expansion of the unknown
functions in an appropriate basis, by diagonalizing the matrix of
the operator ^l2 + 2K^ z in such a basis. Looking at this operator,
one understands how spheroidal orbitals are related to those we
have encountered before. As the distance tends to zero, does
too, therefore the operator becomes ^l2 , which is associated to the
jn; l; mi basis: as ! 0 the jn; f; mi set coincides with the jn; l; mi
set. On the other hand, as the distance between the two centers
goes to innity, the contribution from K^ z becomes dominant, thus
the jn; f; mi coincide with the jn; ; mi. One could work out the
same result by a more complicated study of the asymptotic limits
of the functions jn; f; mi [37]. Considering the eigenvalues (f + 2 ),
what we have just said implies that
lim (f + 2 ) = l(l + 1)
!0
and
lim
!1
f + 2
=
2
(1.52)
35
This conrms the role of spheroidal coordinates as a bridge between
the parabolic and the polar coordinates [17].
Eq. (1.47) allows the determination of the coecients of the
expansion of the basis jn; f; mi by the basis jn; l; mi; we propose
another way to get the jn; f; mi, calculating the coecients of their
expansion in parabolic orbitals jn; ; mi:
Tn;f;m (r) =
n X
jmj 1
= (n jmj 1)
hn;m
f; Un;;m (r)
(1.53)
hn; ; m j ^l2 + 2K^ z j n; 2; mi =
1 2
[n (m + 1)2 ] [n2 (m + 1)]
4
hn; ; m j ^l2 + 2K^ z j n; ; mi =
i
1h 2
(n 1) + (m2 2 ) + 2
2
hn; ; m j ^l2 + 2K^ z j n; +2; mi =
1q 2
[n (m + + 1)2 ] [n2 (m 1)2 ]
4
(1.54)
(1.55)
(1.56)
36
Thus the coecients in Eq. (1.53) satisfy a three terms recurrence:
1q 2
[n (m + 1)2 ] [n2 (m + 1)] hn;m
f; 2 +
4
1 2
1
(n 1) + (m2 2 ) + 2 (f + 2 ) hn;m
f; +
2
2
1q 2
[n (m + + 1)2 ] [n2 (m 1)2 ] hn;m
f;+2 = 0 (1.57)
4
(1.58)
37
= l(l+1) 42
"
1 [n2 l2 ][l2 m2 ]
2l (2l 1)(2l + 1)
(Tn;f;m )[1]
l = un;l;m
1
1) + (m2
2
2 ) +
2
(n
8
(1.60)
2 + m2
1)
(f+ )/(2+n)
38
-2
-2
/4
arctan
Figure 1.2:
/2
39
40
relationship with the distance between the "foci" of the ellipses (see
Fig. 3).
The parametrization in Eq. (1.62) leads to the elliptic harmonics:
(1.63)
^l2
l(l + 1) jl; i = 0
E^ jl; i = 0
(1.64)
(1.65)
41
k =1.00
= /2
y
x
z
k =0.866
= /3
y
x
z
k =0.707
= /4
y
x
z
k =0.500
= /6
y
x
z
k =0.00
= 0.00
Figure 1.3:
y
x
42
2l + 1:
jl; i =
l
X
m= l
xl;m jl; mi
(1.66)
Am 2 xl;m 2 + (Bm
) xl;m + Am xl;m+2 = 0
(1.67)
where
n
o1=2
1
(1 k02 ) (l m)[l2 (m + 1)2 ](l + m + 2)
4
h
i
1
= (1 + k02 ) l(l + 1) m2
2
Am =
Bm
43
relationship between xl;m and xl; m : xl;m = xl; m . From now
on, we will represent the elliptic harmonics by the ket jl; ; ; i.
Indicating the re
ection operators with respect to the planes
yz and xz as X^ and Y^ , one has X^ jl; mi = jl; mi and Y^ jl; mi =
( )m jl; mi. As a consequence, one has: X^ jl; ; ; i = jl; ; ; i
and X^ Y^ jl; ; ; i = jl; ; ; i. The value of the parameter k (or
k0 ) aects , not the other quantum numbers, therefore the symmetry with respect to the re
ection planes is not altered by varying
k (or k0 ). The spheroelliptic orbitals induce irreducible representations of the group D2h [47] [ = ( 1)l ]:
(1.68)
(1.69)
(1.70)
(1.71)
44
with suitable linear combinations of Wigner rotation functions [25]:
i
o1=2
1 nh 2
l (m 1)2 (l m + 2)(l + m)
4
Dml 2;m00 (0; ; ) Dml 2; m00 (0; ; ) +
2 2
2 2
i
1h
2
l(l + 1) m
m002
2
l
l
Dm;m00 (0; ; ) Dm; m00 (0; ; ) +
2 2
2 2
i
o1=2
1 nh 2
l (m + 1)2 (l m)(l + m + 2)
4
Dml +2;m00 (0; ; ) Dml +2; m00 (0; ; ) = 0
2 2
2 2
(1.72)
45
where we have explicitly made the identication with Wigner reduced d-matrix elements for which eq. (1.73) is a well-known property [25]. Comparing Eqs. (1.73) and (1.67) we can see that they
are not equivalent, and consequently one cannot associate a variation of k0 with a rotation of the quantization axis. To conrm this,
we note that for any value of k0 any function jl; ; ; i maintains
its symmetry with respect to the re
ection on xed planes ( and
are not aected by variation of k0 ), and this would not be possible if
the variation of k0 were associated to a rotation of the quantization
axis.
The reasonings about the relationship between spherical and elliptic harmonics can be extended to the corresponding parametrizations. In other words, the Jacobi elliptic functions must tend - as
the modulus takes limiting values - to the ordinary circular functions. Indeed the function sn(; k) can be dened as sn(; k)
sin ' where ' is given by [19]:
Z '
1
d
0 (1 k2 sin2 )1=2
(1.74)
!0 '
For k ! 0, the integrand function tends to one, therefore k=
!0 sin . In Figure 3 we show a representation of the
and sn(; k) k=
spheroelliptic coordinates on the surface of a sphere, for dierent
values of the modulus.
In order to clarify some of the previous concepts, it will be useful
to show an example concerning elliptic harmonics: the case l = 2
will be considered [1]. The elements of the matrix in eq. (1.75) are
46
;k0 = hl = 2; m j^
2
02 l2 jl = 2; mb i, where 2 ma ; mb 2:
Eml=2a ;m
a lx + k ^
y
b
El;k0 =
(1.75)
1
6
02
0
0
(1 + k02 )
0
2 (1 k )
C
B
5
3
0
2
0
2
C
B
0 p 2 (1 + k )
0
(1 k ) p 0
2
C
B
B
6 (1 k02 ) C
02 )
02 ) 6
C
B (1
0
3(1
+
k
0
k
2
2
C
B
3
5
C
B
0
2
0
2
0
(1
k
)
0
(1
+
k
)
0
A
@
2
2
p
6 (1 k02 )
02 )
0
(1
+
k
0
0
2
By appropriately changing the order of rows and columns it is easy
to show that this matrix if made of two blocks, which correspond to
the symmetry classes = 1. Moreover the matrix is centrosymmetric, thus using as a basis the harmonics jl; m; i we get a further
factorization [36]:
J 1El;k0 J =
(1.76)
0
(1 + k02 )
0
3(1 k02 )
0
0
B
C
0
2
B
C
0
(4 + k )
0
0
0
B p
C
B
C
0
2
0
2
B
C
3(1
k
)
0
3(1
+
k
)
0
0
B
C
0
2
B
C
0
0
0
(1
+
k
)
0
@
A
0
0
0
0
(1 + 4k02 )
where
0
1
1 0 0
1 0
B
C
B 0 1
C
0
0
1
B
C
p
1 B
J = p2 BB 0 0 2 0 0 CCC
(1.77)
B 0 1
C
0 0 1 A
@
1 0 0 1 0
Therefore the matrix is factorized in four blocks (three 1x1 and one
2x2) corresponding to the four symmetry classes ( = 1; = 1)
of the elliptic harmonics.
47
Note that one can follow a more straightforward procedure by
directly calculating the matrix of E^ on the basis jl; m; i, instead of
jl; mi: a recurrence relationship identical to Eq.(1.67) is obtained,
except for the matrix element Bm=1; =1 which now reads
1
4 (1
(1.78)
k02 + 1)1=2 ]
p
j2; 1;2; +1; +1i = ( 3=2)(1 k0 2)(k04 k02 + 1) 1=4
[2(k04 k02 + 1)1=2 (1 + k02 )] 1=2
"
#
(1 + k02 ) 2(k04 k02 + 1)1=2
p
(j2; 2i + j2; 2i)
j2; 0i +
02
6(1 k )
3 = 4 + k02
4 = 1 + k02
5 = 1 + 4k02
l(l + 1) m2 = 6 0 = 6
48
. As k02 ! 0, 1 ! 4, the corresponding eigenfunction is indeed a
harmonic dy z (m002 = 4), i.e. the equivalent of dx y , but taking
x instead of z as the quantization axis.
For the case l = 3 (a 7x7 matrix), the -symmetry (even and
odd m) and the -symmetry lead to a factorization into four blocks:
one is 1x1 and three are 2x2. For higher l's, three-dimensional or
larger blocks may result from the factorization. In Table 1 we show
the number of functions in each of the four symmetry classes for
the lower values of l.
2
1.6 Conclusion
In this chapter we have resumed the most important features of
polar, parabolic and spheroidal hydrogenic orbitals, especially emphasizing their use as Sturmian basis sets; we have also studied
their behavior with respect to the inversion and in general to the
symmetry operations of the point group D2h . Spheroidal Sturmians do not belong to the D2h point group, while the other types of
Sturmians are found to induce irreducible representations of that
group, or to do that after an easy transformation, as the one in
Eq. (1.38). Within this context we have given [eq. (1.31)] a general expression for the coecients which conserve the parity in the
transformation between the polar and the parabolic sets. Then we
have presented the features of a lesser known set, the spheroelliptic
one.
49
20
15
10
5
1
0.2
0.4
0.6
Figure 1.4:
0.8
0.2
0.4
0.6
0.8
50
,,
|m >
|lm>
|lm>
|lm>
,
|m >
|m >
|m,, >
|fm>
|lm>
,,
|l >
,
|l >
|lm>
|fm>
|m,>
Figure 1.5:
|l >
|lm>
|fm>
|m >
51
We have paid a great deal of attention to the relationships and
the recoupling coecients among the alternative basis sets. Thus,
now we are able to show (Fig. 5) a complete connection scheme for
alternative hydrogenic Sturmians in conguration space, including
the orthogonal transformations among them. Each connection line
in these gures corresponds to an orthogonal transformation between two alternative basis sets at the extremes of the line. Dotted
lines are associated to the Wigner rotation matrices [18, 24] which
perform the change of the polar axis:
jn; l; mi =
l
X
l
Dm;m
0 (0; ; ) jn; l; m0 i
2 2
m0 = l
jn; f; mi =
nX1
l=jmj
gln;m jn; l; mi
52
in momentum space presented in ref. [49], in order to consider also
elliptic Sturmians and to implement the conservation of parity in
the transformations between any alternative basis sets. We will
compare the results with the ones of the present chapter, nding
analogies and dierences of the treatment of hydrogen atom in the
two reciprocal spaces.
...
=0
=0
fk ) = 0
(1.79)
Therefore the elements of eigenvectors of a tridiagonal matrix enjoy
a three-term recurrence relationship: ci 1;k Fi;i 1 + ci;k (Fi;i fk ) +
ci+1;k Fi;i+1 = 0.
This reasoning can run backward: if we have a transformation
P
between two sets of functions i = k di;k k whose coecients
di;k exhibit a three-term recurrence in k, the passage to the set
i corresponds to the diagonalization of a certain operator whose
matrix calculated on the set k is tridiagonal.
53
The hermiticity requirement for the operators associated to physical observables implies that if the matrix F is real, it has to be
symmetric with respect to the principal diagonal. Therefore if we
have a non-symmetric three term recurrence
ci 1 Xi 1 + ci Yi + ci+1 Zi+1 = 0
(1.80)
Q
Q
Xi 1 i 1 gi 1 + Yi gi + Zi+1 i+1 gi+1 = 0
Qi
Qi
(1.81)
Now let us impose the symmetry with respect to the principal diagonal:
Q
Q
Xi 1 i 1 = Zi i
Qi
Qi 1
Consequently we get the ratio Qi 1 =Qi , which is enough to get an
p
p
explicit symmetrized recurrence: Xi 1 Zigi 1 +Yigi + Xi Zi+1 gi+1 =
0. By such a procedure the Qi are determined to within a constant
that can be xed imposing normalization and a convention for an
overall phase factor.
Bibliography
[1] J. Patera and J. Winternitz. J. Math. Phys., 14:1130, 1973.
54
[2] E.G. Kalnins, W. Miller, and P. Winternitz. S.I.A.M. J. Appl.
Math., 30:630, 1976.
[3] H. Shull and P.-O. Lowdin. J. Chem. Phys., 30:617, 1959.
[4] M. Rotenberg. Ann. Phys. (N.Y.), 19:262, 1962.
[5] M. Rotenberg. Adv. At. Mol. Phys., 6:233, 1970.
[6] J. Avery. Hyperspherical Harmonics, Applications in Quantum
Theory. Kluwer Academic, Dordrecht, The Netherlands, 1989.
[7] J. Avery. Hyperspherical Harmonics and Generalized Sturmians. Kluwer Academic Publishers, Dordrecht, The Netherlands, 2000.
[8] V. Aquilanti, S. Cavalli, C. Coletti, D. Di Domenico, and
G. Grossi. Int. Rev. Phys. Chem., 20(4):673, 2001.
[9] V. Aquilanti and J. Avery. Adv. Quantum Chemistry, 39:71,
2001.
[10] W. Miller Jr. Symmetry and Separation of Variables. AddisonWesley Publishing Company, Reading, Massachusetts, 1977.
[11] W. Miller Jr. Lie Theory and Special Functions. Academic
Press, New York and London, 1968.
[12] V. Aquilanti, S. Cavalli, and M.B. Sevryuk. J. Math. Phys.,
34:3351, 1993.
55
[13] V. Aquilanti, S. Cavalli, and M.B. Sevryuk. J. Math. Phys.,
35:556, 1994.
[14] M. Hamermesh. Group Theory and its Application to Physical
Problems. Dover Publications, Inc., New York, 1989.
[15] C.A.Coulson and A.Joseph. Proc. Phys. Soc. London, 90:887,
1967.
[16] V. Aquilanti, S. Cavalli, and C. Coletti. Chem. Phys., 214:1,
1997.
[17] Stella M. Sung and Dudley R. Herschbach. J. Chem. Phys.,
95:7437, 1991.
[18] R. N. Zare. Angular Momentum. Wiley & Sons, New York,
1988.
[19] M. Abramowitz and I.A. Stegun. Handbook of Mathematical
Function. Dover, New York, 1964.
[20] P. M. Morse and H. Feshbach. Methods of Theoretical Physics.
McGraw-Hill, New York, 1953.
[21] V. Aquilanti, S. Cavalli, C. Coletti, and G. Grossi. Chem.
Phys., 209:405, 1996.
[22] B. R. Judd. Angular Momentum Theory for Diatomic
Molecules. Academic Press, New York, 1975.
56
[23] L.D. Landau and E.M. Lifshitz. Quantum Mechanics. Nonrelativistic Theory. Addison-Wesley Publishing Co., Reading
Mass., 1958.
[24] M.E. Rose. Elementary Theory of Angular Momentum. Wiley,
New York, 1957.
[25] D.A. Varshalovich, A.N. Moskalev, and V.K. Khersonskii.
Quantum Theory of Angular Momentum. World Scientic,
Singapore, 1988.
[26] D. Park. Z. Phys., 36:155, 1960.
[27] T. Regge. Nuovo Cimento, 10:544, 1958.
[28] V. Aquilanti, L. Beneventi, G. Grossi, and F. Vecchiocattivi.
J. Chem. Phys., 89:751, 1988.
[29] V. Aquilanti and G. Grossi. J. Chem. Phys., 73:1165, 1980.
[30] V. Aquilanti, S. Cavalli, and G. Grossi. Z. Phys. D., 36:215,
1996.
[31] V. Aquilanti, S. Cavalli, and G. Grossi. Theor. Chim. Acta,
79:283, 1991.
[32] V. Aquilanti and S. Cavalli. Few-Body Systems. Suppl.6: 573,
1992.
[33] K. Schulten and R.G. Gordon. Comp. Phys. Comm., 11:269,
1976.
57
[34] K. Schulten and R.G. Gordon. J. Math. Phys., 16:1961, 1975.
[35] V. Aquilanti, S. Cavalli, and D. De Fazio. J. Phys. Chem.,
99:15694, 1995.
[36] A.R. Collar. Quart. Journ. Mech. and Applied Math., 15:265,
1962.
[37] C.A.Coulson and P.D.Robinson. Proc. Phys. Soc. London,
71:815, 1958.
[38] M. Aubert, N. Bessis, and G. Bessis. Phys. Rev. A, 10:51,
1974.
[39] M. Aubert, N. Bessis, and G. Bessis. Phys.Rev. A, 10:61, 1974.
[40] P.D. Robinson. Proc. Phys. Soc. London, 71:828, 1958.
[41] T. Levitina and E. J. Brandas. Int. J. Quantum Chem., 65:601,
1997.
[42] T. Levitina, E. J. Brandas, and B. Larsson. Int. J. Quantum
Chem., 85(4-5):392, 2001.
[43] I. V. Komarov, L. I. Ponomarev, and S. Yu. Slavyanov.
Spheroidal and Coulomb Spheroidal Functions. Nauka,
Moskow, 1976.
[44] I.S. Gradshteyn and I.M. Ryzhik. Table of Integrals, Series
and Products. Academic Press, New York, 1980.
58
[45] C.A. Coulson and A. Joseph. Int. J. Quantum Chem., 1:337,
1967.
[46] E. T. Whittaker and G. N. Watson. A Course of Modern
Analysis. Cambridge University Press, London, 1935.
[47] J. Patera and P. Winternitz. Journal of Chemical Physics,
65:2725, 1976.
[48] L. Pauling and E.B. Wilson Jr. Introduction to Quantum Mechanics. McGraw-Hill, New York and London, 1935.
[49] V. Aquilanti, S. Cavalli, and C. Coletti. Phys. Rev. Lett.,
80:3209, 1998.
Chapter 2
Hydrogenic Orbitals in Momentum
Space and Hyperspherical Harmonics.
Elliptic Sturmian Basis Sets.
Summary
Momentum space hydrogenic orbitals can be regarded as orthonormal and complete Sturmian basis sets, and can be explicitly given in terms of (hyper)-spherical harmonics on the fourdimensional hypersphere S 3 . Among the alternative coordinate systems which allow separation of variables, the usual ones involving
parametrizations of the sphere S 3 by circular functions correspond
to canonical subgroup reduction chains; we also investigate harmonic "elliptic" sets (as e.g. obtained by parametrizations in terms
of Jacobi elliptic functions). In this chapter we list the canonical
hydrogenic Sturmian sets and the orthogonal transformations connecting them. The latter enjoy very useful three-term recurrence
59
60
relationships which allow their ecient calculations even for large
strings. We also consider modications needed when the conservation of the symmetry of Sturmians with respect to parity. Finally
we discuss some properties of elliptic hydrogenic Sturmians and
their relations with canonical Sturmians. Since elliptic Sturmians
cannot be expressed in closed form, it is important to nd expansions in a suitable basis set and to calculate the transformation
coecients. We derive three-term recursion relationships fullled
by the coecients of the transformation between elliptic Sturmians
and canonical Sturmians. A concluding discussion on the connections between conguration space and momentum space hydrogenic
Sturmians completes this chapter.
2.1 Introduction
The investigation of the non-relativistic Schrodinger equation for
hydrogenic atoms in momentum space has great relevance not only
for the obvious importance of the physical problem, but also because it allows the construction of basis sets enjoying dierent symmetry properties useful in atomic and molecular structure calculations and for the description of atoms in elds.
Hydrogenic orbitals in momentum space, rst obtained by Podolski and Pauling [1] by direct Fourier transform, can indeed be identied, within a normalization factor, with hyperspherical harmonics of a 4-dimensional sphere, as shown by Fock [2] through his
61
famous stereographic projection. Modernly they can be referred
to as momentum space Sturmians [3{7], being the counterpart of
the so called "natural spin" [8] or conguration space Sturmian
orbitals [9, 10], which are nding increasing applications in atomic
and molecular quantum chemistry. For recent reviews, see Refs. [11]
and [12]. The reciprocity existing between conguration and momentum space has to be applied with care when trying to extend all
the symmetry properties which can be found in momentum space,
to the corresponding conguration space Sturmians. The symmetry
and completeness properties of these sets make them in fact adapt
to solve quantum mechanical problems where the hyperspherical
symmetry of the kinetic energy operator is broken by the interaction potential, but the corresponding perturbation matrix elements
can be worked out explicitly, as in the case of Coulomb interactions.
In this chapter, after explaining the connection between the
hydrogen atom and the four-dimensional hypersphere (section 2),
we review and update the classication of hydrogenic Sturmians
in momentum space given in ref. [13] (section 3), requesting the
conservation of parity (section 4). In addition elliptic Sturmians in
momentum space are introduced (section V), in which the modulus
of Jacobi elliptic functions can be seen as an additional `degree of
freedom', making them more
exible with respect to the canonical
Sturmians. As also stressed in the concluding Sec. 6, the aim of
this chapter has been to show that they are less `intractable' than
previously believed [14].
62
2.2 Background
In 1935 Fock obtained the wavefunctions of hydrogen atom (p)
in momentum space, by solving the Schrodinger Equation related
by Fourier transform to the one in conguration space:
1 Z
(p2 + p20 ) (p) = 2
(p0 )
jp p0j2 dp
(2.1)
(2.2)
pz = p cos #
where p is the modulus of p and # and ' its polar angles, is projected onto a 4-dimensional hypersphere of unit radius:
2p p
x = 2 0 x2 = sin sin # cos '
p0 + p
2p p
y = 2 0 y2 = sin sin # sin '
p0 + p
2p p
z = 2 0 z2 = sin cos #
p0 + p
p2 p2
w = 02 2 = cos
p0 + p
(2.3)
63
where 0 . Eq. 2.3 exhibits the relationship between the
Euclidean space (px ; py ; pz ) and the S 3 surface (x2 + y 2 + z 2 + w2 =
1), and also gives an explicit parametrization in polar coordinates
(; #; ') of the S 3 hypersphere. Fig. 1 illustrates Fock's stereographic projection, but in order to make it readable we have drawn
it considering one dimension lower, i.e. projecting a 2-dimensional
plane on the 3-dimensional sphere S 2 , instead of the 3-dimensional
space on the 4-dimensional hypersphere S 3 .
The eigenfunction for the hydrogen atom in momentum space
is [6, 7, 12]:
n;l;m (p) =
4p50=2
Y (; #; ')
(p20 + p2 )2 n;l;m
(2.4)
64
u2
,
u1
u3
p1
p2
Figure 2.1:
65
the degeneracy existing for a given n gives rise to n2 Sturmians
which belong to the same irreducible representation of O(4).
The hydrogenic hamiltonian can be expressed as a function of
^
the orbital angular momentum ^l and of the Runge-Lenz vector A
^ [15, 18]:
in its reduced form K
1
^ 2 + 1) 1
H^ = (^l2 + K
2
where
(2.5)
1 ^
^l + np0r =p0
l
p
^
p
^
(2.6)
2
r
^ can be identied with the
The Cartesian components of ^l and K
generators of four-dimensional rotations, in virtue of the isomorphism of the hydrogen atom with the sphere S 3 . Let us consider
the Euclidean four-dimensional space: there are six possible distinct
bidimensional planes identied by the various couples of Cartesian
axes; by indicating one of those planes as ef , a rotation of an angle
upon it is performed by the operator eiJ^ef , where J^ef is one of
the six generators of O(4) [17, 19]:
K^ = A^ =p0 =
@
1
J^ef = ue
i
@uf
@
uf
@ue
(2.7)
fg J^eh
eh J^fg + eg J^fh
(2.8)
66
[19]:
@2
1 @ d 1@
= d 1
r
2
@ue r @r
@r
1 X ^2
J
r2 f>e ef
(2.10)
A d-dimensional hyperspherical harmonic solves the eigenvalue equation on the d-dimensional sphere S d 1 parametrized by the set of
angles
d 1 :
X
b>a
J^ab2 Yj;(
d 1 ) = j (j + d 2) Yj;(
d 1 )
(2.11)
b>a
J^ab2 Yj;(
3 ) = j (j + 2) Yj;(
3 )
(2.12)
where denotes two quantum numbers. Use of eq. (2.9) shows that
^ 2, and therefore
the operator in eq. (2.12) coincides with ^l2 + K
momentum space Sturmians (which, as we have seen, essentially
coincide with four-dimensional harmonics) can be identied with
basis sets for irreducible representations of the rotational group
O(4), as suggested by eq. (2.4). The hyperangular momentum
quantum number j = n 1 labels the energy levels.
That of eq. (2.3) is only one of the possible parametrizations
separating the Hamiltonian for the motion on S 3 (for it, = l; m):
67
starting from other suitable parametrizations alternative momentum space orbitals can be worked out. A complete classication of
the bases of O(4), including symmetry properties and connections
among them will be given in the following sections.
68
0
S (,,)
m
l
(a)
n-1
0 =2
3
S (,,)
z x
n-1
Figure 2.2:
(b)
69
the two schemes for the construction of S 3 , and shows connections
between Cartesian and hyperspherical coordinates and the representation of the corresponding hyperspherical harmonics according
to the tree method. This graphical method, a full account of which
can be found in refs. [20{23], permits a visualization not only for the
connection between coordinates, but also for the rotation operators
for which the harmonics are eigenfunctions.
The following paragraphs will deal with the classication and
description of the various sets corresponding to the dierent parametrizations. Obviously the choice of a particular basis set is completely
pointless when the symmetry of the problem is the original S 3
one (as in the case of hydrogenoid atoms), but it becomes crucial when such a symmetry is broken, and yet the perturbation
hamiltonian can still be written in terms of the generators of the
four-dimensional rotation group, as in the case for instance for hydrogenoid atoms in electric and/or magnetic elds. Important in
quantum chemistry are molecular orbitals which describe symmetry
breakings occurring when atoms join to form molecules.
2.3.1
The harmonics in g. 2a correspond, as sketched before, to a subgroup reduction chain O(4) O(3) O(2), where a fourth dimension is added to the Euclidean tridimensional space [x; y; z ] whose
generators are ^lx , ^ly and ^lz , or according to the previous identi-
70
cation, J^12 , J^23 , J^13 (see Eq. 2.9). Explicitly these harmonics can
be written as the product between the usual tridimensional spherical harmonic Yl;m (#; '), eigenfunctions of the quadratic "orbital
angular momentum" operator ^l2 = ^lx2 + ^ly2 + ^lz2 with eigenvalues
l(l + 1), and a properly normalized Gegenbauer polynomial [24] of
the additional variable cos (Eq. 2.3):
(2.13)
(2.14)
The harmonics Yn;l;m(; #; ') are also eigenfunctions of ^lz with eigenvalue m; this corresponds to the choice of the (xy ) plane where the
rotation subgroup O(2) acts and thus of z as the polar axis. Continuing to identify O(3) as acting on the tridimensional space [x; y; z ]
- the fourth dimension being not involved yet - when the (yz ) or
(xz ) planes are chosen for the O(2) subspaces one has the harmonics Yn;l;m0 (; #; '0 ) and Yn;l;m00 (; #; '00 ), eigenfunctions of ^ly (y
polar axis) and ^lx (x polar axis) respectively. Interchanges among
the polar axes, illustrated for example in g. 3, can be performed
by means of a Wigner rotation matrix:
m0
l
Dm;m
0 (0; =2; =2) Yn;l;m (; #; ')
(2.15)
Let us now consider a three-dimensional subspace where the rotation group O(3) acts, and let w be one of the axes: two of the
71
(a)
n-1
m
Wigner rotation matrix
n-1
Figure 2.3:
Representation of the change of quantization axis for the spherical basis sets from m (z axis) to m0 (y axis), corresponding to Eq. (2.15).
(b)
72
harmonics, where 0 and 00 denote the corresponding eigenvalues.
The connection between these bases and the Zeeman basis (compare
with the illustration in g. 3 of the case of the spherical basis) is
given by a Wigner rotation matrix, similar to eq. (2.15). Beside the
Zeeman basis set (and those basis sets with a dierent projection
axis, i.e. Yn;;0 (0 ; 0 ; 0) and Yn;;00 (0 ; 00 ; 00)), there are six other
bases of the same kind which can be identied: three of them are
eigenfunctions of the quadratic operator K^ x2 + ^ly2 + K^ z2 (with eigenvalue 0 (0 +1)) and of K^ x , ^ly and K^ z respectively (with eigenvalues
00 , m0 and ), while the last three bases are diagonalized by the
quadratic operator ^lx2 + K^ y2 + K^ z2 and by the operator ^lx , K^ y and
K^ z respectively (with eigenvalues m00 , 0 and ). A summary of the
twelve bases of O(4) corresponding to asymmetric parametrizations
in given in Table 1.
Orthogonal transformations which lead to a change in the tridimensional angular momentum quantum number (transformations
among l; ; 0 ; 00 ) can be performed by means of the Z matrices introduced in [13] (see also [28]). For example the orthogonal transformation between theP spherical (eq. 2.4) and the Zeeman basis
n;m
sets [Yn;l;m(; #; ') = Zl;
Yn;;m(0 ; ; ')] is depicted in g. 4.
In ref. [13] it was shown that the harmonic superposition between
the two bases, which had previously been dened indirectly through
the steps: Spherical basis ! Stark basis ! Zeeman basis and calculated as a sum on two vector coupling coecients [25,29] (see Eq.
2.22), can be written as a single sum of the Racah type:
n;m = ( )l+ [C (l)C ()]
Zl;
X
1
2
(2.16)
73
d = 3-subspace
d = 2-subspace
(x; y ) ^lz
(y; z ) ^lx
(z; x) ^ly
basis set
(x; y ) ^lz
(y; w) K^ y
(x; w) K^ x
jn; ; mi
jn; ; 0i
jn; ; 00i
(z; x) ^ly
(z; w) K^ z
(x; w) K^ x
jn; 0; m0 i
jn; 0; i
jn; 0; 00i
(y; z ) ^lx
(y; w) K^ z
(y; w) K^ y
jn; l; mi
jn; l; m00i
jn; l; m0i
74
(a)
/2
Z matrix
n-1
(b)
n-1
Figure 2.4:
where
( n+2k 1 + p(k))!( k+2m q (k))!( k 2m q (k))! ( n 2 k + p(k))
C (k)= n+k+2
( 2 + p(k)) ( m+2k+1 + q (k)) ( k m2 +1 + q (k))( n 2k 1 p(k))!
n;m
is zero
p(k) = 1+( 41) and q (k) = ( 1) 4 1 . Note that Zl;
n;m
n; m
when n + l + + m is even and shows the symmetries Zl;
= Zl;
n;m
n;m
and Zl;
= Z;l
. The sum in 2.16 [28] is a hypergeometric function
4 F3 of unit argument and can be connected with Racah polynomials
[30, 31], although it cannot be reduced to the ordinary Racah's
or 6-j coecient which performs angular momentum recoupling.
Indeed, like a Racah's recoupling coecient it is orthogonal with
respect to summation on two angular momentum quantum numbers
(l and ), but contains a projection quantum number. As shown
n;m
in Ref. [13], the Z;l
coecient can be compactly written as a 6-j
symbol extended to allow not only multiples of 1/2, as ordinary
k n
k m
75
vector recoupling coecients, but also multiples of 1/4:
n;m = (
Zl;
l+
2
+1+p(l)+p() (l + 1 )( + 1 )
2
2
The symbol
1
4
1
4
p()
p(l)
n 1
2
m 1
2
l
1
4
1
4
(2.17)
Symmetric Parametrizations
We will deal now with the symmetric parametrizations, corresponding to the subgroup reduction chain O(4) O(2) O(2). The best
known of these basis sets is that corresponding through a Fourier
transform to the Sturmian orbitals written in parabolic coordinates
in conguration space (see chapter 1). Due to its importance in the
treatment of the hydrogen atom in an electric eld, we call the corresponding harmonic set as the Stark basis. For its use for building
molecular orbitals and multidimensional expansions, see [3,4,12,32].
This parametrization is illustrated in g. 2b and can be written as
(2.18)
1
2
76
simultaneously diagonalize ^lz and K^ z (with eigenvalues m and ),
where we give to the z axis the role of a privileged direction in
physical space. This explains why the Hamiltonian for a hydrogen
atom in a weak electric eld directed along the z axis is still diagonal
in this basis. Similarly other two basis sets of this kind can be
constructed when the x and y axes are taken as quantization axes
(see Table 2). Thus we can have a basis diagonalizing ^lx and K^ x ,
and another one diagonalizing ^ly and K^ y , leading to a total of three
dierent basis sets corresponding to symmetric parametrizations.
The orthogonal transformation between spherical (g. 2a) and
Stark basis (g. 2b), analogously to what is found in conguration
space (see Chapter 1), is given by a Clebsch-Gordan coecient [33]:
cn;m
l; n;l;m (p)
(2.20)
1
m 1
m+
; (n 1);
jl; mi
cn;m
l; = h 2 (n 1);
2 2
2
(2.21)
n;;m (p) =
( )
1+m
2
O(4)
We have just shown that the number of distinct canonical basis sets
for S 3 amounts to 15. Actually the total number of solutions to
77
Laplace equation on S 3 (harmonics) is 120, however due to the fact
that for the identication of a plane or of a 3-dimensional space it is
only necessary to specify the involved axes and not their sequence,
to dierent systems of hyperspherical coordinates, identifying the
same plane, there correspond the same harmonics (or harmonics
diering for a phase factor only, i.e. still eigenfunctions of the
same rotation operator), so the number of dierent bases reduces
to 15. Ref. [13,34] provides a discussion on the classication of the
120 coordinate systems and of how this reduction is achieved. The
15 basis sets can be depicted together with their connections by
the graph in Fig. 5. This is a projective plane RP 2 representation,
which in turn can be obtained from a graph, associated to the
icosahedral group, illustrating the 120 coordinate systems (see ref.
[13]). The connections between dierent basis sets, which are placed
at the vertices of 10 triangles and 6 pentagons, are identied by
various types of segments: (i) vector coupling coecients (eq. 2.20)
are drawn by solid lines, (ii) rotation matrix elements (eq. 2.15) by
dotted lines and (iii) Z coecients (eq. 2.17) by zig-zag lines. Fig. 5
is also an "abacus" to obtain relationships among ordinary elements
of angular momentum algebra, augmented by the Z coecients:
following the sides of the plane gures one can write interesting
sum rules. For example following the side of the triangle where m
n;m
is conserved Zl;
can be written as a sum involving two ClebschGordan coecients
n;m =
Zl;
X
( ) h
(2.22)
n 1+m n 1 m
1+m n 1 m
;
;
; jl; 0ih
;
;
; j; 0i
2
2
2
2
2
2
2
2
(the phase = n 1 2m + l + is an integer: if it were omitted, the sum would be l ). We recognize on the right hand side
the transformation coecient used for the denition of the Zeeman
basis [25]. An alternative sum rule can be obtained from g. 5,
78
following the sides of a pentagon:
n;m X
0
dlm;m0 ( )Zl;
= ( 1)m+ dm;0 ( )
2
2
(2.23)
0
0
0
h n 12+ m ; 2 ; n 12
m0 0
n
; ; l; 0ih
1 + 0 m0 n 1 0 m0
;
;
; ; ; 0i
2
2
2
2
[(n 1)2
l2 ][(n
1)2
2 ][(n
(l + 1)2 ]
1)2
n 2;m +
( + 1)2 ] Zl;
[(n 1)2
2 l + l + 2 l l2 + l2 + 2 l2 +
n;m
n2 (1 + + 2 + l + l2 2m2 n2 ) + 2m2 ] Zl;
+
n[
n
q
[(n + 1)2
1q
[(n + 1)2
l2 ][(n + 1)2
2 ]
(2.24)
n+2;m = 0
( + 1)2 ] Zl;
2
2l + 3 u
t (l
2
m2 )[(l
1)2
m2 ](n2 l2 )[n2
(2l 3)
(l
1)2 ]
Zln;m2;
79
000
111
000000000
111111111
000
111
000000000
111111111
000
111
00000000
11111111
00000000
11111111
000000000
111111111
000
111
11111111
00000000
00000000
11111111
000000000
111111111
00000000
11111111
000000000
111111111
00000000
11111111
000000000
111111111
00000000
11111111
000000
111111
000000000
111111111
00000000
11111111
000000000
111111111
00000000
11111111
000000
111111
000000000
111111111
00000000
11111111
00000000
11111111
000000
111111
000000000
111111111
00000000
11111111
00000000
11111111
000000
111111
000000000
111111111
00000000
11111111
00000000
11111111
000000
111111
000000000
111111111
00000000
11111111
000000000
111111111
000000
111111
000000000
111111111
00000000
11111111
000000000
111111111
000000
111111
000000000
111111111
00000
11111
00000000
11111111
000000000
111111111
000000
111111
000000000
111111111
0000011111111
11111
00000000
000000000
111111111
000000
111111
000000000
111111111
00000
11111
00000000
11111111
000000000
111111111
000000
111111
000000000
111111111
0000011111111
11111
00000000
000
111
000000000
111111111
000000
111111
000000000
111111111
0000011111111
11111
00000000
0000
1111
000
111
000000000
111111111
000000
111111
000
111
00000
11111
00000000
11111111
0000
1111
000
111
000000000
111111111
0000000
1111111
000
111
00000
11111
0000
1111
000
111
000000000
111111111
0000000
1111111
000
111
00000
11111
0000
1111
0000000001111
111111111
0000000
1111111
000
111
00000
11111
0000
000000000
111111111
0000000
1111111
000
111
00000
11111
0000
1111
000000000
111111111
0000000
1111111
000
00000
11111
0000111
1111
0000000001111
111111111
0000000
1111111
000
111
00000
11111
0000
000000000
111111111
0000000
1111111
000
111
00000
11111
0000
0000000001111
111111111
0000000
1111111
000
00000
11111
0000111
1111
0000000001111
111111111
0
1
0000000
1111111
000
111
00000
11111
0000
000000000
111111111
0
1
0000000
1111111
000
111
00000
11111
0000111
1111
000000000
111111111
0
1
0000000
1111111
000
00000
11111
0000
1111
0000000001111
111111111
0
1
0000000
1111111
000
111
00000
11111
0000
000000000
111111111
0
1
0000000
1111111
000
111
0000111
1111
000000
111111
000
111
000000000
111111111
0000000000
1111111111
000
000000000000
111111111111
0000
1111
000000
111111
000
111
00000000
11111111
0000000001111
111111111
0000000000
1111111111
000
111
000000000000
111111111111
0000
000000
111111
000
111
00000000
11111111
0000000000
1111111111
000
111
000000000000
111111111111
0000
1111
000000
111111
000
111
00000000
11111111
0000
1111
0000000000
1111111111
000
111
000000000000
111111111111
000000
111111
000
111
00000000
11111111
0000
1111
0000000000
1111111111
000000000000
111111111111
000000
111111
000
111
00000000
11111111
0000
1111
0000000000
1111111111
000000000000
111111111111
000000
111111
000
111
00000000
11111111
0000
1111
0000000000
1111111111
000000000000
111111111111
000000
111111
000
111
00000000
11111111
0000
1111
0000
1111
0000000000
1111111111
000000000000
111111111111
000000
111111
000
111
00000000
11111111
0000
1111
0000
1111
0000000000
1111111111
000000000000
111111111111
000000
111111
000
111
00000000
11111111
0000
1111
0000
1111
0000000000
1111111111
000000000000
111111111111
000000
111111
000
111
00000000
11111111
0000
1111
0000
1111
0000000000
1111111111
000000000000
111111111111
000000
111111
000
111
00000000
11111111
0000
1111
0000
1111
0000000000
1111111111
000000000000
111111111111
000000
111111
000
111
00000000
11111111
0000
1111
0000
1111
0000000000
1111111111
000000000000
111111111111
000000
111111
000
111
00000000
11111111
0000
1111
0000
1111
000000
111111
000000000000
111111111111
000000
111111
00000000
11111111
0000
1111
0000
1111
000000
111111
000000000000
111111111111
000000
111111
000000000
111111111
00000000
11111111
0000
1111
00000000
11111111
0000
1111
000000
111111
000000
111111
000000000
111111111
00000000
11111111
0000
1111
00000000
11111111
0000
1111
000000
111111
000000000
111111111
00000000
11111111
0000
1111
00000000
11111111
0000
1111
000000
111111
000000000
111111111
00000000111111111
11111111
0000
1111
00000000
11111111
0000
1111
000000
111111
000000000
00000000
11111111
0000
1111
00000000
11111111
0000
1111
000000
111111
000000000
111111111
00000000
11111111
00000000
11111111
000
111
000000
111111
000000000
111111111
00000000
11111111
00000000
11111111
000000000
111111111
000
111
000000
111111
000000000
111111111
00000000
11111111
00000000
11111111
000000000
111111111
000
111
000000
111111
000000000
111111111
00000000
11111111
00000000
11111111
000000000
111111111
000
111
000000
111111
000000000
111111111
00000000
11111111
000000000
111111111
000000
111111
000000000
111111111
00000000
11111111
000000000
111111111
000000
111111
000000000
111111111
00000000
11111111
000000000
111111111
000000
111111
000000000
111111111
00000000
11111111
000000000
111111111
000
111
000000
111111
000000000
111111111
00000000
11111111
000000000
111111111
000
111
00000000
11111111
000000000
111111111
000000000
111111111
000
111
00000000
11111111
000000000
111111111
000000000
111111111
00000000
11111111
000000000
111111111
00000000
11111111
| l m>
| >
| m>
| >
| l m>
|l m>
| m>
|m>
|m>
|> |
> |>
|>
|m>
| m>
Figure 2.5:
80
1
pn + 2
2l
1 [(l + 1)2
m2 ][(l + 2)2
m2 ][n2
2l + 5
(l + 1)2 ][n2
Zln;m
+2; = 0
(l + 2)2 ]
(2.25)
or m:
m + 1q
( m + 2)(l m + 2)[(m 1)2 2 ]
2
q
n;m 2
(l + m)( + m)(l + m)[(m 1)2 l2 ] Zl;
+ m[
2
l + l + l2
l2 + l2 + l2 2
n;m
+m2 (1 + + 2 + l + l2 m2 2n2 ) + 2n2 ] Zl;
m 1q
( m)(l m)[(m + 1)2 2 ]
(2.26)
2
q
[(m + 1)2 l2 ]( + m + 2)(l + m + 2) Zl;n;m+2 = 0
Note that these relations are extremely useful for calculation of full
strings of coecients.
81
physical systems and thus it is important to have Sturmian bases
with denite parity.
2.4.1
(2.27)
p1
[ n;;m(p) ( )m n;
p)]
;m (
(2.28)
and the conservation of parity in the passage to the set n;l;m(p)
involves the parity-conserving Clebsch-Gordan coecients (pCG):
hn; l; m j n; ; m; i = (
1+m
2
Cl;n;m;
jj = ( )
1+m
2
1 + ( )l n;m
cl;jj
2(1 + ;0 )
(2.29)
82
which have been dened in the previous chapter Eq. (1.31), Sec.
1.3.
2.4.2
83
The corresponding momentum space Sturmian can be expressed as
(see eq. 2.13):
n;;m (p) =
4p5=2
( )n 1 jmj ({)+m jmj 2 0 2 2 Bn; cos 0 Cn+11 (sin 0 ) Y;m(; ')
(p0 + p )
(2.32)
where Bn; had been dened in eq. (2.14). Taking into account
that inversion changes the sign of p = (px; py ; pz ) but leaves p2
unchanged, from eqs. (2.30) and (2.31) one sees that such an operation corresponds to the transformation of angles (0 ; ; ') into
( 0 ; ; ' + ). Considering that
1 C +1 (sin 0 )
n 1
eim('+) = ( )m eim'
P^ n;;m (p) = ( )n
1+m
n;;m(p)
(2.33)
Thus we have shown that the basis n;;m (p) is the set of eigenfunctions of the operator which performs the inversion of p, and the
same proof applies to the bases n;0 ;m0 (p) and n;00 ;m00 (p), thus
the three matrices Z which - as in g. 4 - perform the transformation to the functions n;l;m(p), n;l;m0 (p) and n;l;m00 (p) should
be factorized in two blocks, in order to conserve the parity. These
conclusions are consistent with the observation that the matrix eln;m
ements Zl;
are zero when n + l + + m is even [13, 36].
84
The transformation between Stark and Zeeman basis sets (eq.
2.20 with instead of l) involves the pCG coecients, which have
been amended by requesting the conservation of parity [compare
with eq. (2.29)]. One has to consider that the Zeeman basis set has
parity ( )n 1+m :
p)
n;jj;m(
n X
jmj 1
( )
jj=0 or 1
C;n;m;
jj =
1+m
2
C;n;m;
jj n;;m (p)
(2.34)
(2.35)
85
p2 p2
w = 02 2 = cos 0 sin 0 sin '00
p0 + p
thus one can write
p2 + p2
cos 0 sin cos '00
px = 0
2p0
p2 + p2
py = 0
cos 0 cos
(2.36)
2p0
p2 + p2
sin 0
pz = 0
2p0
Therefore the inversion of p in the tridimensional space corresponds
to the transformation of angles (0 ; ; '00 ) ! ( 0 ; ; '00 ).
The expression of the Sturmian is, apart from a normalization factor
which is not needed for symmetry considerations:
p50=2
0 C +1 (sin 0 ) P j00 j (cos ) ei00 '00
cos
n 1
2
(p0 + p2 )2
(2.37)
Because of the change in the sign of the argument of the exponential
function as a consequence of the inversion, it is evident that the
Sturmian in eq. (2.37) cannot be eigenfunction of P^ . Applying
the projection procedure (p) = p12 [ n;;00 (p) n;;00 ( p)] one
can work out:
n;;00 (p) =
n;;j00 j (p) =
p1
n;;00 (p) ( )n
00 1
p)
n;; 00 (
(2.38)
86
parity components of a Sturmian which was originally eigenfunction
for one of the K's Cartesian components, one has to combine two
Sturmians having opposite values of (or 0 or 00 ). Thus, the
orthogonal transformation
n;;m (p) =
00
00
;
Dm;
j00 j (0; =2; =2) n;;00 (p)
where
;
Dm;
j00 j (0; =2; =2) =
1+
( )n 1+m
q
Dm;00 (0; =2; =2) (2.39)
2(1 + 00 ;0 )
The form of eq. (2.39) is due to the fact that matrix elements
00 (0; =2; =2) diering only in the sign of 00 can only have
Dm;
equal or opposite values.
In summary, the basis sets which have (or 0 or 00 ) as a
quantum number can be turned into "parity Sturmians" performing
a linear combination with the Sturmians labeled by (or 0 or
00 ). These "parity Sturmians" will no longer be labeled by
(or 0 or 00 ), but by jj (or j0 j or j00 j) plus , the eigenvalue
of the inversion operator P^ . As a consequence, matrix elements of
orthogonal transformations where jj (or j0j or j00 j) is conserved as
a quantum number have to be premultiplied by the suitable parity
factor, which appears in eqs. (2.29),(2.34) and (2.39).
87
88
ships fullled by the transformation coecients among canonical
and elliptic Sturmians will be given.
2.5.1
w = cos
(2.40)
89
l, (k)
(a)
n-1
(b)
n-1, b(k)
m
,
(c)
n-1, q(k)
Figure 2.6:
90
2.5.2
(2.41)
(
cos m'
The corresponding hyper-harmonics are given as L()L0( )
sin m'
0
where L() and L ( ) are solutions of equations related to the socalled associated Lame equation [15]: these solutions are not available in closed form, but information on a certain basis set and on
its connections with the others can be obtained by studying the
corresponding commuting operators.
By setting F^j = 2i (^lj + K^ j ) and G^ j = 2i (^lj K^ j ) (j = x; y; z )
such operators are [15]:
^ 2 + G^ 2)
C^ = 2(F
(2.42)
!
k2 + 1 ^ ^
^
^
^
^
^
B = Fx Gx + Fy Gy + 2
FG
k 1 z z
R^ = (F^z + G^ z )2
(2.43)
(2.44)
91
the meaning of C^ and R^ is clear: they are absolutely equivalent to
^ 2 and ^lz2. Now we have to explain the
the familiar operators ^l2 + K
meaning of the operator B^ . Eq. (2.43) can be rewritten as
F^ G^ =
k2 1 z z
1 (^lz + K^ z )(^lz K^ z )
^ G^
=F
k2 1
2
^
^
The product F G can be expressed as:
^2 ^
F^ G^ = 12 (F^ + G^ )2 12 (F^ 2 + G^ 2) = l2 + C4
Therefore the elliptic cylindrical harmonics of type I diagonalize
the operator
^l2 C^ 1 (^lz + K^ z )(^lz K^ z )
+
2 4
k2 1
2
Multiplying by -2 and expanding, the latter is found to be equivalent to
^l2 + 1 2 K^ z2 + 2 1 ^lz2 C^
(2.45)
1 k
k 1
2
Consequently, the basis can be characterized by the set of operators
^l2 + K^ 2, ^l2 + 1 1k K^ z2 and ^lz2. In the limit k ! 0, the second one
tends to ^lx2 +^ly2 +^lz2 + K^ z2 , which by an appropriate linear combination
with C^ and ^lz2 gives K^ x2 + K^ y2 + ^lz2 : this is tantamount to say that
the elliptic cylindrical harmonics of type I tend, in this limit, to
the Zeeman set Yn;;m . In the opposite limit, the operator tends to
K^ z2 , thus the set tends to that of denite parity Stark harmonics
Yn; jj;m.
2
92
The elliptic cylindrical coordinates of type II are (see g. 6c):
z = sn(; k) dn(; k0 );
The operators which characterize the corresponding set of hyperharmonics are the ones in eqs. (2.42) and (2.44), plus [15]
(2.48)
By some passages similar to the ones we made for the elliptic base
of type I, Q^ can be written as:
^l2 k2 ^ ^ ^ ^ C^
+ (l + Kz )(lz Kz ) +
Q^ =
2 2 z
4
A more perspective equivalent expression for this operator is
^l2 + k2K^ z2 k2^lz2 C^
(2.49)
2
Therefore this basis can be characterized by the set of operators
^l2 + K^ 2, ^l2 + k2K^ z2 and ^lz2. As k ! 0 it goes into the spherical basis,
since ^l2 + k2 K^ z2 ! ^l2 , while as k ! 1 one has the operator ^l2 + K^ z2 ,
which by linear combination with C^ and R^ is seen to correspond to
the operator which characterizes the Zeeman basis.
Now we can insert the elliptic cylindrical bases of type I and
II in a scheme (gure 7) where their relationships with the other
bases are shown. The basis I, as the modulus k varies, changes
continuously from Yn;;m to Yn; jj;m, and the basis II from Yn;l;m to
Yn;;m .
93
y w
n1, q(k)
n1, b(k)
Ell. Cylindrical
type II
k
1
lm
nm
Z l
z
Ell. Cylindrical
type I
k
k
C ||
m
n1
n1
z
n1
Figure 2.7:
n1
|| m
nm
Relationships among elliptic and canonical Sturmians. For limiting values of the modulus k , the elliptic Sturmians coincide with the canonical
ones, i.e. those corresponding to binary trees.
94
It is possible to get the unknown elliptic cylindrical eigenfunctions by calculating the coecients of their expansion in suitable
basis sets. As we diagonalize the matrix of ^l2 + k2 K^ z2 on the basis
^ 2 and ^lz2, therefore
Yn;l;m, the eigenvectors diagonalize also ^l2 + K
they are elliptic cylindrical functions of type II. By calculating the
matrix of the operator ^l2 + k2 K^ z2 on the basis Yn;l;m we can prove
P
that the coecients in Yn;q;m
= l bn;m;
q;l Yn;l;m - where Yn;q;m is the
elliptic cylindrical basis set of type II and q is the eigenvalue of
^l2 + k2K^ z2 - are related by the three-term recursion relationship
r
q bn;m;
q;l +
m ] n;m;
bq;l+2
2
=0
(2.50)
95
culate the overlaps, i.e. the transformation coecients, between
Yn;b;m
(the elliptic cylindrical basis set of type I, where b is the
eigenvalue of the operator ^l2 + 1 1k K^ z ) and Yn; jj;m. Putting
P
Yn;b;m = jj an;m;
b;jj Yn;jj;m one has, by calculating the matrix of
^l2 + 1 1k K^ z2 on the basis Yn; jj;m:
2
1 2
[n (m + 1)2 ] [n2 (m + 1)2 ] an;m;
b;j 2j +
4
1 2
1
1
(n 1) + (m2 2 ) +
2 b an;m;
b;jj +
2
2
2
1 k
1q 2
[n (m + + 1)2 ] [n2 (m 1)2 ] an;m;
b;j+2j = 0 (2.51)
4
According to the scheme of gure 7, the latter recurrence relationship is diagonal as k ! 1, while as k ! 0 eq. (2.51), it coincides
with the recursion in for the coecients C;n;m;
jj , and the eigenvalue b becomes (n2 1) ( + 1) + m2 . In g. 8 we show the
correlation diagrams for the separation constants b and q of the elliptic cylindrical orbitals of type I and II in the case n = 3; m = 0.
q
96
4
Elliptic Cylindrical Basis, Type II
q
n+k
0
2
Arctan k
Figure 2.8:
/4
b
2
n+1/(1-k )
0
/2
Arctan 1/(1-k )
97
connections between them. Within the framework of information
already available in the literature [4, 13] we have introduced explicitly the requirement of the conservation of parity with respect
to inversion and have obtained three-term recurrence relationships
for the transformation coecients, including the Z -matrix elements
recently dened [25, 26, 39]. Our detailed investigation has lead to
three-term recurrence relationships which provide a simple calculation scheme for the development of cylindrical elliptic orbitals of
rst and second type both in the familiar spherical and in the Stark
orbitals. This also has allowed to elucidate the relationships with
respect to the `canonical' bases.
Finally, amplifying the results of the previous chapter which
focused on orbitals in conguration space, we can discuss analogies and dierences of the obtained basis sets in the two reciprocal
spaces. Of the four types of orbitals in conguration space, three
of them (polar, parabolic, spheroelliptic) have a counterpart in momentum space (spherical, Stark, spheroelliptic) respectively. The
fourth type of orbitals (spheroidal), characterized by being eigenfunction of the operator ^l2 + 2K^ z , has no counterpart in momentum space. On the other hand, three types of orbitals among those
available in momentum space (Zeeman, elliptic cylindrical of type
I, elliptic cylindrical of type II) characterized respectively by being
eigenfunctions of the operators K^ x2 + K^ y2 + ^lz2 , ^l2 + 1=(1 k2 )K^ z2
and ^l2 + k2 K^ z2 , have no counterpart in conguration space. The
reason for this "missing correspondence" (one manifestation being
98
the intractability of spheroidal orbitals according to Coulson [14])
^ operator: according
can be attributed to the dierent role of the K
to its denition (Eq. 2.6), it can be seen that it is a second order
dierential operator in conguration space, while it acts as a rst
order dierential operator on eigenfunctions in momentum space.
This is at variance with the role of ^l, which is of the rst order in
both spaces. Since the conguration space Schrodinger equation
is of the second order, as it is its counterpart describing motion
on S 3 , separability of variables implies the action of second order
operators [15, 40]. On the other hand in conguration space there
^ 2 (or any of its cartesian
are no diagonal operators which contain K
components).
The fact that each alternative type of orbitals in momentum
space provides eigenfunctions of cartesian components of ^l2 and
K^ 2 or of their linear combinations, also implies that these orbital
sets induce irreducible representations in the group D2h [15]: the
correspondence holds for the reciprocal conguration space only
in the case of polar, parabolic and spheroelliptic orbitals, which
also induce irreducible representations in D2h , while the spheroidal
orbitals do not.
A detailed description of Sturmian basis sets and of their counterpart in momentum space is extremely important for the implementation of the remarkable plane wave expansion:
exp(ip r) = (2 )3=2
p2 + p2
un;l;m(r) 0 2 n;l;m (p)
2p0
n;l;m
X
(2.52)
99
which is seen to be essentially an expansion in four-dimensional
spherical harmonics and Sturmian polar orbitals (their explicit expression can be found in the previous chapter. Eq. (2.52) is a key
formula in momentum space quantum chemistry and was introduced by Shibuya and Wulfman [41] in 1965 who used it to expand
one-electron molecular orbitals and to obtain secular equations for
multicenter problems. A detailed mathematical analysis for such
an expansion can be found in ref. [42].
Previously, we had introduced an extension of such a formulation to Stark Sturmians [3], which has proved to have competitive convergence properties when applied to the calculation of the
ground state energy of the hydrogen molecular ion. This research
has emphasized the importance of exploiting the diverse alternative
separable coordinate systems. The extension of such expansions to
the dierent types orbitals treated in this chapter seems extremely
promising and stimulating, not only in the three-dimensional case,
but also in the d-dimensional wave expansion, as already done [4]
for spherical and Stark Sturmians. The solution of many-body
Coulomb problems would in fact require multidimensional plane
wave expansions [12].
Bibliography
[1] B. Podolsky and L. Pauling.
[2] V. Fock.
100
[3] V. Aquilanti, S. Cavalli, C. Coletti, and G. Grossi.
Phys., 209:405, 1996.
[4] V. Aquilanti, S. Cavalli, and C. Coletti.
1997.
Chem.
[9] M. Rotenberg.
[10] M. Rotenberg.
101
[14] C.A.Coulson and A.Joseph.
1967.
[18] L.D. Landau and E.M. Lifshitz. Quantum Mechanics. Nonrelativistic Theory. Addison-Wesley Publishing Co., Reading
Mass., 1958.
[19] B. R. Judd. Angular Momentum Theory
Molecules. Academic Press, New York, 1975.
[20] Yu.F. Smirnov and K.V. Shitikova.
1976.
for Diatomic
102
editors, New Methods in
1996.
of Mathematical
[30] A.F. Nikiforov, S.K. Suslov, and V.B. Uvarov. Classical Orthogonal Polynomials of a Discrete Variable. Springer-Verlag,
Berlin, 1991.
[31] V. Aquilanti, S. Cavalli, and C. Coletti.
344:587, 2001.
103
[34] V. Aquilanti and C. Coletti. Chem. Phys. Lett., 344:601, 2001.
[35] V. Aquilanti, S. Cavalli, and G. Grossi.
1996.
85:1362, 1986.
[43] A. A. Izmest'ev, G. S. Pogosyan, A. N. Sissakian, and P. Winternitz. J. Math. Phys., 40:1549, 1999.
[44] E. G. Kalnins and Jr. W. Miller.
1986.
104
d = 2-subspace d = 2-subspace
(y; z ) ^lx
(x; w) K^ x
basis set
jn; 00; mi
(z; x) ^ly
(y; w) K^ y
jn; 0; mi
(x; y ) ^lz
(z; w) K^ z
jn; ; mi
Chapter 3
Many-electron one-center systems
Summary
This chapter deals with the application of polar Sturmian orbitals to the solution of the Schrodinger equation for free atoms
and atoms in external elds. We dene a N -electronic generalized
Sturmian basis set as a product of one-electron Sturmian orbitals.
Expanding the electronic wavefunction by such a generalized Sturmian basis set, we obtain a secular problem in which the matrix
elements do not contain any kinetic energy operator. The solution of the secular problem yields accurate values for the energies
of electronic ground and excited states. In the Sturmian method
framework, the eect of strong external elds can be taken into
account with little additional eort; we provide the formalism for
the treatment of external perturbation and show the results of a
preliminary calculation for static polarizability of Lithium.
105
106
3.1 Introduction
Chapters 1 and 2 contain a deep insight into the properties of the
dierent species of hydrogenic Sturmian orbitals. In this chapter
we will see the application of polar Sturmian orbitals (Eq. 1.7),
to the calculations of physical properties for isolated atoms. Illustrating the approach to atomic structure as a many-electron
one-nucleus problem, we discuss here the procedure of building a
polyelectronic basis set as a product of "monoelectronic" Sturmian
orbitals [1{3, 7, 9, 11, 12]: an alternative method, which denes N electronic Sturmians as solutions of the 3N -dimensional hydrogen
atom can be elegantly formulated and also has a momentum space
counterpart in terms of hyperspherical harmonics [4{6], but still
lacks of developments in the algebraic and computational implementation in order to be computationally competitive.
Sturmian approach is quite dierent from the procedure used
commonly in modern ab initio quantum chemistry:
The one-electron orbitals are Sturmian orbitals. Their analytic expression is known a priori, except the exponent of
the radial part, whose numerical value is obtained after the
calculation of the properties of the system.
(In the most common ab initio procedures for quantum mechanical calculations, Gaussian functions are used as basis
set. An Hartree-Fock iterative calculation is performed to
combine those basis functions into one-electron orbital).
107
where
1 2
p2
r
+ V0 (x) + V 0 (x) + 0 (x) = 0
2
2
N
X
r2 = r2j
j =1
V0 (x) =
and
V 0 (x) =
(3.2)
N
X
Z
j =1 rj
(3.3)
N
X
1
i>j rij
(3.1)
(3.4)
p
108
The wavefunction is expanded by a linear combination of Slater
determinants
(x) =
where
C (x)
(x) =
(3.5)
x
x
un
un
( 1)
N ;lN ;mN ;msN
N ;lN ;mN ;msN ( 2 )
un
;l ;mN ;msN ( N )
..
.
N N
..
.
(3.6)
The one-electron functions un;l; m; ms (xj ) are the polar Sturmian orbitals given in Eq.(1.7), multiplied by the spin function (ms =
1=2); x stands for fx1; x2 ; : : : ; xN g and fxj g = frj ; #j ; 'j g.
Our basis set is not completely dened yet, because we have
to establish a rule for assigning a value to the exponents kj . We
demand our basis functions (x) to be a solution to the dierential
Equation [1, 8, 9, 11, 12]:
"
1 2
p20
r + V0(x) + 2 (x) = 0
2
(3.7)
(3.8)
n1 k1 = n2 k2 = ::: = nN kN = Z
(3.9)
and
109
Eqs.(3.8) and (3.9) imply that the exponents' value is:
kj =
p0
1
1
nj n + n + + n1N
q
2
1
2
2
(3.10)
2E
(3.11)
110
Eq.(3.1) yields:
X
"
p2
1 2
r
+ V0 (x) + V 0 (x) + 0 C (x) = 0
2
2
(3.12)
Then, taking into account that the functions (x) obey Eq.(3.7),
we can eliminate the kinetic energy operator from Eq.(3.12):
X
(3.13)
2E 0 ;
C = 0 (3.14)
[T 0 ;
p0 0 ; ] C = 0
(3.18)
11=2
X
2E
p
1
T00 ; =
0 ; = 0 0 ; = = @ kj2 A
p0
j
11=2
1A
0 ; = Z @
2
j nj
(3.19)
X
0 ;
111
The integral T00 ; doesn't depend on p0 's value, and this is true also
for the electronic repulsion integral T0 0 ; : in the radial part of the
integral, one can perform a change of integration variable from r to
p0 r, so that p0 's value is not needed to solve the integral. In order to
calculate energies and wavefunction for the electronic states of an
isolated atom, one has just to compute and diagonalize the matrix
T 0 ; .
X
V 00 (x) = E rj cos #j
(3.20)
j =1
p0 0 ; C = 0
(3.21)
where
T000 ;
dx 0 (x)
N
X
j =1
p0 rj cos #j (x)
(3.22)
(3.23)
2
p0
112
The value of integral (3.22) is not a function of p0 . The procedure
for solving Eq.(3.21) is the following:
1. Establish a value for ;
2. Compute the matrix elements;
3. Diagonalize the matrix, getting p0 ;
4. The electric eld intensity is E = p20 ;
5. The energy of the atom in the electric eld is p20 =2.
(3.24)
where the sum is extendend to all possible permutations of electrons, and p is the permutations parity.
113
3.4.1
Overlap Integrals
f1 (1) f2 (1)
f1 (2) f2 (2)
1
1
(x) = p F (x) = p
..
..
N!
N!
.
.
f1 (N ) f2 (N )
0 i between two N -
fN (1)
fN (2)
..
.
..
.
fN (N )
(3.25)
and 0 (x) = p1N ! G(x) similarly dened. The overlap can be expressed as:
h j 0 i = N1 ! hF j Gi
1 ^
hA[f1 (1) : : : fN (N )] j A^[g1 (1) : : : gN (N )]i
N!
= hf1 (1) : : : fN (N ) j A^[g1 (1) : : : gN (N )]i
hf jg i hf jg i hf jg i
1
1
1
2
1
N
hf2 jg1 i hf2 jg2 i hf2 jgN i
= ..
..
..
..
.
.
.
.
hfN jg1 i hfN jg2 i hfN jgN i
un0 ;l0 ;m0 (r; #; ') un;l;m(r; #; ')r2 dr sin #d# d' =
= 23+2l (k
k
0 )l+3=2
"
(n l 1)!(n0 l 1)!
2n[(n + l)!]3 2n0 [(n0 + l)!]3
#1
2
(3.26)
114
Z
The result of the angular integral is nl;n0 l0 , while the radial integral can be solved considering the explicit expression for Laguerre
polynomials:
Z
(n0 + l)!
n0X
l 1(
1)k
k!
k=0
nX
l 1 n0X
l 1
nX
l 1(
j =0
1)j
j!
n+l
n l j 1
n0 + l
n0 l j 1
(2k r)j 5
(2k r)k 5 dr =
( 2k )j ( 2k0 )k Z 1
dr exp (k +k0 )r r2l+2+j +k
j
!
k
!
0
j =0 k=0
The latter radial integral is easy to solve:
Z 1
(2l + j + k + 2)!
dr exp (k +k0 )r r2l+2+j +k =
(k + k0 )2l+j +k+3
0
Considering Eq.(3.10), one would expect the result of Eq.(3.26) to
be a function of p0 , but a deeper analysis indicates that a change of
variable could be performed inside the integral, from r to p0 r; this
implies the result to be independent from p0 's numerical value.
(n+l)!(n0 +l)!
3.4.2
115
We want to obtain a general expression for the matrix element
h j V j 0 i. Considering the formulas given above, it is straightforward to show that
(3.28)
F=
G=
N
X
( 1)i+1 fi (1)FiN 1
(3.29)
(3.30)
i=1
N
X
j =1
hF j v(1) j Gi =
N X
N
X
i=1 j =1
h j V j 0 i =
N X
N
X
i=1 j =1
(3.32)
where jSij j is the minor obtained from the determinant of the overlap matrix by deleting the ith row and the j th column.
If the operator in Eq.(3.27) is the Coulomb potential for an atom
embedded in a static electric eld (see section 3.3), then the matrix
116
element in Eq.(3.32) corresponds to the integral
Z
un0 ;l0 ;m0 (r; #; ') p0 r cos # un;l;m(r; #; ') r2 dr sin #d# d' =
Z
#1
"
(n l 1)!(n0 l0 1)!
0
2n[(n + l)!]3 2n0 [(n0 + l0 )!]3
0
0
exp (k +k0 )r rl+l +3 L2nl+1l 1 (2k r)L2nl0 +1l0 1 (2k0 r)dr
0
0
= 23+l+l kl+3=2 kl 0+3=2 p
Z
Yn0 ;l0 (#; ') cos #Yn;l (#; ') sin #d#d'
(3.33)
The radial part of this integral is very similar to the one in Eq.(3.26):
Z
0
0
exp (k +k0 )r rl+l +3 L2nl+1l 1 (2k r)L2nl0 +1l0 1 (2k0 r)dr =
2
0
exp (k +k0 )r rl+l +3 4(n + l)!
2
4(n0 + l0 )!
n0 X
l0 1 (
k=0
1)k
k!
nX
l 1
( 1)j
j!
j =0
n0 + l 0
n0 l 0 j
(2k
r )j 5
n+l
n l j
(2k
r)k 5 dr
nX
l 1 n0 X
l0 1
( 2k )j ( 2k0 )k Z 1
0
dr exp (k +k0 )r rl+l +3+j +k
j !k!
0
j =0 k=0
The solution of the latter integral is:
(n+l)!(n0 +l0 )!
3.4.3
(l + l0 + j + k + 3)!
0
dr exp (k +k0 )r rl+l +3+j +k =
(k + k0 )l+l0 +j +k+4
0
1
In this paragraph we are going to demonstrate generalized SlaterCondon rule for matrix elements between determinants involving
117
two-electron operators. The typical example of a two-electron operator is the Coulomb repulsion among N electrons, corresponding
to the sum of N (N 1)=2 terms 1=rij . If we dene the potential as
V0 =
1
j>i rij
(3.34)
1)
1
h
Fj
j Gi
2)!
r
12
(3.35)
We now expand the determinants F and G in terms of their double
minors:
F=
N X
N
X
fj (1)fi (2)]FijN 2
(3.36)
(3.37)
i=1 j =i+1
and
G=
N X
N
X
k=1 l=k+1
In Eq.(3.36), FijN 2 is the double minor obtained from the determinant F by deleting the rst and second rows and the ith and j th
columns. Considering that A2N 2 = (N 2)!AN 2 , and combining
Eqs.(3.35), (3.36) and (3.37), we get
N
X
h j V 0 j 0 i =
N X
N X
N
X
where
Cij ;kl =
1
[g (1)g (2)
r1 2 k l
118
and jSij ;klj is the determinant of the matrix obtained by deleting
ith and j th row, and kth and lth column from overlap matrix S .
From Eq.(3.39), we understand that in order to calculate the
matrix element given in Eq.(3.35), we have to solve integrals of the
tipe:
Z
1
dx1 dx2 un ;l ;m (x1 )un ;l ;m (x1 ) un ;l ;m (x2 )un ;l ;m (x2 )
r12
(3.40)
The operator can be expanded as [10]:
1
l
1 X
r<
4 X
1
=
r12 r> l=0 m= l r>
!l
1
Y (
)Y (
)
2l + 1 l;m 1 l;m 2
(3.41)
where Yl;m (
i ) are normalized spherical harmonics and
i = f#i ; 'i g.
Inserting the latter expansion into Eq.(3.40), and separating radial
and angular integrals (un;l;m(xj ) = Rn;l (rj ) Yl;m (
j )), we obtain :
1 Z
(
)
4
d
1 Yl;m (
1 )Yl ;m (
1 )Yl;m
1
l;m 2l + 1
X
(3.42)
d
2 Yl;m (
2 )Yl ;m (
2 )Yl;m (
2 )
3
r< l 1
r> r>
The integral of the product of three spherical harmonics is:
dr1 r12 Rn ;l (r1 ) Rn ;l (r1 )
1
v
u
u
t
(
)Y 0 0 (
)Y 00 00 (
) =
d
Yl;m
l ;m
l ;m
(2l0 + 1)(2l00 + 1)
hl; 0jl0; 0; l00; 0ihl0; m0; l00 ; m00jl; mi
4 (2l + 1)
(3.43)
119
So we can use Eq.(3.43) to solve the angular integral in Eq.(3.42),
that becomes:
4
l=max(jl
l j;jl l
min(l X
+l ;l +l )
2
(2l1 + 1)(2l + 1)
1
hl2 ; 0jl1 ; 0; l; 0ihl1 ; m1; l; mjl2 ; m2i
2
l
+
1
4
(2
l
+
1)
2
j)
(3.44)
(2l4 + 1)(2l + 1)
hl3 ; 0jl4 ; 0; l; 0ihl4 ; m4 ; l; mjl3 ; m3 iRn ;l
4(2l3 + 1)
1
;n ;l ;n ;l ;n ;l ;l
2
Z 1
Z r1
+1
1
0
rl
dr
Z 1
+2
l
dr2 r
2 Rn3 ;l3 (r2 )Rn4 ;l4 (r2 ) + r1
l
(3.45)
1
dr2 r
2
Rn ;l
4
Y
k
l3 1
3X
c
l1 1
1X
2f
1;3 )a
a!
=0
2+l+lX
3 +l4 +c+d
=0
1 l+lX
3 +l4 +c+d
i
=0
c
)
lk
c!
=0
2f
3;1
=1
(2fk )
1 + l1
n1
1 a
n
l
(nk
lk
nk (nk
l4 1
i1=2 n4X
lk )!
3 + l3
l3
c
2;4
2f
b
)
b!
=0
4 ;2 ) d
2f
d!
=0
h
l2 1
n2X
1)!
4 + l4
n4
l4
d
2 + l2
l2
b
1
(3.46)
(2 + l + l3 + l4 + c + d)!(1
l + a + b + l1 + l2 )!
3;1 + f4;2 )3+c+d+l+l3 +l4 (f1;3 + f2;4 )2 l+a+b+l1 +l2
(f
(2 + l + l3 + l4 + c + d)!(1
l + a + b + l1 + l2 + i)!
3 1 + f4;2 )3+c+d+l+l3 +l4 i (f1;3 + f2;4 + f4;2 + f3;1 )2 l+a+b+l1 +l2 +i
i!(f ;
3 4
j !(f3;1 + f4;2 )2 + 3 + 4 + +
(1
1 + l2 + a + b)!
4 2 + f3;1 )3+l+j+l1 +l2 +a+b
l + l + l + c + d)!(2 + l + j + L
l l l
c d j (f
; +f ; +f ;
13
24
fp;q
np n1q + n1p
2
(3.47)
120
1 2 ~
p~2
r
+ V0 (x) + 0 ~ (x) = 0
2
2
(3.48)
"
p~20
2
(3.49)
"
T; 0
p~
S 0
p~0 ; 0 + 0 S; 0 + ; E C = 0
2
p~0
(3.50)
The symbol S; 0 stands for the overlap between basis functions:
S; 0
(3.51)
121
Eq.(3.50) is a generalized eigenvalue problem, where the eigenvalue
is the energy E . We have to investigate which is the best value for
p~0 , and whether it coincides with p0 .
In the particular case of a single basis function ~ (x) (a trivial
1x1 matrix) our problem becomes:
p~0 E
+ =0
2 p~0
T;
(3.52)
p~2
E= 0
2
p~0 T;
(3.53)
p~
E
2 + p~
T1;1
p~
T1;2 + 2 S1;2 +
0
S;
p~
1 2
0
T2;2
p~
2 + p~
0
=0
(3.54)
122
If we assign p~0 a value of 2.47 a.u., Eq.(3.54) gives E = 2:8557
a.u., while exploiting Eq.(3.18) one obtains p0 = 2:389 a.u. and
thus E = 2:8533 atomic units. This means that in general p0 is
not the best choice for the denition of the energy of the basis set.
3.6 Examples
When calculations over free atoms are performed using Eq.(3.18)
convergence rate toward the ground state energy is good, and moreover the upper eigenvalues give very good approximations to the
excited state energies. We performed a calculation over Helium (1 S
states), using a basis set made of 1 S bielectronic Sturmian functions with quantum numbers up to n1;max = 4; l1;max = 2; n2;max =
10; l2;max = 2. Accurate 1 S states for helium are available since
long time [13], and continue to be a benchmark for alternative basis sets in variational calculations [14,15]. The ground state energy
we have obtained is 2.902690 a.u., to be compared with the best
variational value 2.903724377 a.u. (error = 0:036%). Considering
also upper eigenvalues, we have been able to reproduce carefully
by our ab initio calculation the Helium 1 S electronic transition
energies between the ground state and several excited states (Table
3.1). In the literature there are some recent examples of accurate
calculations on Helium excited states [16, 17] performed with various methods, but they seem to be not suitable for extensions to
more complex atoms.
123
1 S Levels
1s2
1s 2s
1s 3s
1s 4s
1s 5s
1s 6s
1s 7s
(a.u)
-2.902690
-2.147417
-2.061402
-2.033607
-2.021181
-2.014546
-2.010585
Transition energies
1s ns 1s2 (cm 1 )
165763
184641
190742
193469
194925
195794
Spectra Database
asd).
Experimental(b)
166277.547
184641.369
190741.633
193468.726
194924.965
195794.328
(http://physics.nist.gov/cgi-
124
literature [18].
Bibliography
[1] O. Goscinski, Preliminary Research Report No. 217, Quantum
Chemistry group, Uppsala University (1968).
[2] J. Avery,
673 (1992).
[8] J. Avery, Hyperspherical Harmonics and Generalized Sturmians. Kluwer Academic Publishers, Dordrecht, The Netherlands
(2000).
[9] V. Aquilanti and J. Avery,
125
71
(2001).
486 (1994).
J. Phys. B, 31,
[18] R.W. Molof, H.L. Schwartz, T.H. Miller and B. Bederson Phys.
Rev. A, 10, 1131, 1974.
126
Chapter 4
Sturmian approach to one-electron
many-center systems: integrals and
iteration schemes
Summary
128
4.1 Introduction
Shibuya and Wulfman rst proposed the use of suitably normalized
Coulomb orbitals as basis sets for one-electron molecules in quantum chemistry, exploiting the momentum space approach [1], where
these orbitals are related to hyperspherical harmonics. Later, this
matter was further developed by other researchers [2, 3, 8, 10{13,
19,23{27] who when working in conguration space generally refer
to these orbitals as Sturmians. In the literature there are also recent and extensive reviews and books [4, 20, 30]. By means of this
method, the Schrodinger equation is trasformed into an algebraic
eigenvalue problem, whose matrix elements are linear combinations
of nuclear attraction integrals (NAIs) involving Sturmians. Actually, it is possible to formulate the algebraic problem according to
dierent levels of iterations, which are not equivalent when using
a nite-size basis set, as necessary when truncation is applied in
explicit calculations. The most extensive results on the electronic
energy of H+2 were obtained by Koga and Matsuhashi, who demonstrated the extreme accuracy of the Sturmian method [13].
In Sec. 4.2 we reconsider the basic integrals S and W [1], for
polar and parabolic Sturmians; in Sec. 4.3 we show examples of calculations for H+2 in order to demonstrate that the Sturmian method
yields satisfactory accuracy also in the case of a small-size basis set,
and then we report a quantitative comparison among the performances of the dierent iterations of the wave equation.
129
1 2 1 2
r + 2 p0
2
np0
f (x) = 0
r
(4.1)
nZ
1
dx f0 (x) f (x) = 0 ;
p0
jxj
(4.2)
nZ
R) = p0 dx unlm(x) jx1j un0l0m0 (x
n0 l0 m0 (
Snlm
R)
(4.3)
130
(R is the distance between the two centers) can be obtained working
in momentum space and exploiting angular momentum quantum
theory [1]. The following form for the integral is essentially due to
Dunlap [2]; we have rederived it according to our phase conventions
for the orbitals [21]:
n0 l0 m0 (
Snlm
R) = (
0
)l l+M 2
p30
8 0
n
>
>
< 2
v
u
1
n0 1
2
l0
N;L
n 1
2
n 1
N 1 >
=
2 >
N 1
[
>
>
;
R)
NLM (
L)(N + L + 1)
1u
t (N
N +1;LM (R)
2
N (N + 1)
v
u
u
t
1 (N + L)(N L
2
N (N 1)
Here,
M
8
9
1)
N 1;LM (R)5
>>=
> is a 9j symbol.
>
>
: >
;
>
>
<
00 0
1Z
1
dx unlm (x)
p0
jx Rj un0l0 m0 (x)
(4.5)
jx
1 r
1 X
<
=
x0 j r> l=0 r>
!l
4 X 0
Y (^x )Ylm (^x)
2l + 1 m lm
(4.6)
131
n0 l0 m0 (R):
we have obtained the following expression for the integral Wnlm
"
0 0
0 0
R) = 2 (n + l)!(n + lnn)!(0n(2l +l 1)(21)!(l00n+ 1)l
n0 l0 m0 (
Wnlm
l00
#
1)!(2l0 + 1) 1=2
(4.7)
nX
l 1
( 1)j
j =0 j !(2l + j + 1)!(n l
n0 X
l0 1
( 1)k
0
0 l0
k=0 k !(2l + k + 1)!(n
1)!
(l + l0 + l00 + j + k + 2)!
1)!
(2s)l00 +1
2l00
a l00 1
2s (l + l0 + l00 + j + k + 2)! X (2s)
+ e 2s
a
!
a=0
00 +j +k+1 "
l+l0 lX
b=0
(l + l0
l00 + j + k + 1)!
b!
W =
00
=
(l + l0 + l00 + j + k + 2)
b+l00
(2
s
)
;
(b + 2l00 + 1)!
00 1
00
S 00 S0
n
(4.8)
132
involving those functions. In Ref. [21], the question of the evalun0 0 m0 was tackled expanding parabolic functions by the
ation of Snm
polar ones, and then exploiting the result in Eq.(4.4). Here we show
n0 0 m0 is studied directly
the result that is obtained when integral Snm
in momentum space, according to Shibuya and Wulfman's spirit:
n m (R) =
Snm
0 0 0
1
nZ 3
dx Unm (x) Un0 0 m0 (x
p0
jxj
R) =
(4.9)
!
nn0 1=2 X n 1 + m n0 1 0 + m0 n00 1 00 + m00
=2
h 2 ; 2 ; 2 ; 2 j 2 ; 2 i
p30
n00
00
1 4u
00 )2 m002
t (n + 1
Un00 +1;00 1;m00 (R)
4
n00 (n00 + 1)
v
u
u
t
v
u
u
t
v
u
u
t
(n00
1 + 00 )2 m002
Un00 1;00 1;m00 (R)+
n00 (n00 1)
(n00
39
=
1 00 )2 m002
5
00
00
00
U
(
R
)
n
1
;
+1
;m
;
n00 (n00 1)
133
in terms of a Jacobi polynomial (Ref. [17], page 138); then the
integration over is performed by Eq. 7.391(2) of Ref. [18], yielding
a generalized hypergeometric series 3 F2 (1); the latter is replaced by
a Clebsch-Gordan coecient, according to Eq.(22) at page 241 of
Ref. [17]. The nal expression is easily programmable in Fortran
or in Mathematica language:
Z
0 0 0R
n m
Wnm
n+
nn0
= 2
0
n + 0
n m
2
X
a
0 +0
2
X
c
1)
c!(c +
m0
)!
^
Yl00 ;m00 R
1)
"
0 +0 1
2
00
(2l
00 +1
(2s)l
( +
00
+ 2)!
a
;
b+c
2s
+ 00 +2
X
l
=0
( +
00
i!
+ 0
2
b+d+ m m
00
00
m0
i1=2
1)
n m
1)
)!
+ 0 +m00 ! a + c + m+m0
2
2
( + l00 + 1)!
( + 1)(
d!(d +
a+c+ m m
+ 1)
+ 0 +m00 !
2
+m
0 1
=0
hl00 ; m00
2
X
(4.10)
0
n
b!(b + m)!
0 0
=0
2
X
+ 1
2
"
n++m
n m
a!(a + m)!
( ) =
0
n + 0 + m0
b+d+ m m
x jx Rj Un00 m0 x
n m
(
=0
00
0 1
X
=0
dx Unm (
0
n
) =
0 0 1
2
00
00 #1=2
)!
a
;
#
+ 2)!
(2s)
00
b+c
d+m
2
X
l00 +1
00
s l e 2s
j =0
(2 )
(
00
j!
)
+ 1)!
(2s)
134
Table 4.1: Nuclear attraction integrals W involving parabolic Sturmians, Eq.(4.10).
n; ; m n0 ; 0 ; m0
1,0,0
1,0,0
1,0,0
2,-1,0
2,1,0
2,-1,0
1
n0 0 m0 (R) = 1 R dx3 U
Wnm
nm (x) jx Rj Un0 0 m0 (x)
p
1 e s (1 + s)
s
s
0
1,0,0
2,-1,0
2,1,0
2,-1,0
2,1,0
2,1,0
2 s+e
(2+5
p s+6s +4s )
2 s 2s
2 s e p(2+3s+2s )
2 2s
3+3s+2s e s (3+9s+14s +13s +8s +4s )
2s
3 3s+2s e s (3+3s+2s +s )
2s
3+e s (3+6s+6s +4s +2s )
2s
2s
"
p2
1 2
r
+ v (x) + 0 c f = 0
2
2
(4.11)
v (x) =
va (x) =
Za
jx Xaj
(4.12)
135
Exploiting Eq.(4.1), and dividing Eq.(4.11) by p0 , one obtains the
generalized eigenvalue problem [4, 9]:
X
(A; 0
p0 C; 0 ) c = 0
(4.13)
where
1Z 3
dx f 0 (x)v (x)f (x)
p0
n Z 3 va (x)
C; 0 =
dx f 0 (x)
f (x)
p0
Za
A; 0 =
(4.14)
(4.15)
Z1
0
;0 + Z2 W (R)
0
n
Z1 Z2 0
+
S (R)
Af1;g;f2;0 g =
n n0
Cf1;g;f1;0 g = ;0
0
Cf1;g;f2;0 g = S (R)
Af1;g;f1;0 g =
(4.16)
(4.17)
(4.18)
(4.19)
136
-0.6
-0.8
{1s}
-1
{1s, 2s}
-1.2
E ( = - p0 / 2 )
-1.4
-1.6
-1.8
-2
R ( = s / p0)
Figure 4.1: Energy of the electron ground state in H+2 . The circles
stand for the 10-gure accuracy data by Koga et al. [13]. The three
curves represents the results of Eq.(4.13) for three dierent polar basis
sets: f1sg; f1s; 2sg; f1s; 2s; 2pz g.
137
As it is clearly demonstrated in Ref. [4], there is an interesting
relationship between matrix elements A; 0 and C; 0 , which could
not be gured out at a rst glance but is connected to Eq.(4.8):
A; 0 =
00
Z 00
C; 00 00 C 0 ; 00
n
(4.20)
This allows a simpler form for obtaining eigenvalues and eigenvectors of the one-electron system [4, 9]
X
0s
@
Za Za0
A 0
nn0 ;
; 0 p20 A c = 0
(4.21)
An even simpler form was given by Novosadov [23] (see also [2]):
X
0s
@
Za Za0
C 0
nn0 ;
; 0 p0 A c = 0
(4.22)
138
eigenvalue problems and generalized eigenvalues problems, we have
used Lapack libraries written in Fortran, available on the Internet.
Fig.(4.2) clearly demonstrates that Eq.(4.13) is much more convenient than the others. In order to obtain more accurate results
than those reported in Fig.(4.2) for Eq.(4.13) by enlarging the basis set, we encountered problems of a computational nature, due to
excessive closeness to zero of some elements of the C matrix. These
diculties can be interpreted as an indication of the tendency of
the basis set towards overcompleteness [16], when its dimension is
increased: in other words, since C matrix can be viewed as an overlap matrix in a Sobolev space [33], zero eigenvalues are signatures
for linear dependence among functions in the basis set.
The most direct way for avoiding such a problem was proposed
by Lowdin and others [14, 15], who suggested to proceed with an
expansion of the molecular orbital in terms of a complete set centered on a single point in the space, which will be the midpoint in
the H+2 case. It is interesting to explore the suggestion, inserting
such a single-point expansion in Eq.(4.11): one obtains the secular
problem:
Xh
nl
00
0 0i
n l m (s=2) k n l c
2Wnlm
nl
nlm; = 0
(4.23)
139
-1
1.0010
1.0010
-2
first iteration
1.0010
1.0010
-3
second iteration
-4
1.0010
1.0010
-5
third iteration
-6
50
100
140
given in Fig.(4.3). As the internuclear distance increases, a larger
and larger basis set would be necessary to approximate correctly
the H+2 wavefunction, and one should face the problem of computn0 l0 m0 matrix elements for high quantum numbers, but the
ing Wnlm
accuracy obtained at such an elementary level is most encouraging.
141
-0.8
-1.2
E = - k /2 (a.u.)
-1
-1.4
-1.6
-1.8
-2
2
R = s / k (a.u.)
Figure 4.3: The circles stand for the 10-gure accuracy data by Koga
et al. [13]. The three curves represents the results of Eq.(4.23) for three
dierent basis sets: fnmax ; lmax g = f3; 2g; f6; 5g; f8; 6g.
142
Bibliography
[1] T. Shibuya and C.E. Wulfman,
(1965).
[2] Brett I. Dunlap
for Diatomic
[4] J. Avery, Hyperspherical Harmonics and Generalized Sturmians. Kluwer Academic Publishers, Dordrecht, The Netherlands
(2000).
[5] J.F. Rico, R. Lopez and G. Ramrez,
(1988).
Comp.
Comp. Phys.
143
(1955).
Chem.
144
[24] Ch. Duchon, M.Cl. Dumont-Lepage and J.P. Gazeau, J. Phys.
A, 15, 1227 (1982).
[25] Ch. Duchon, M.Cl. Dumont-Lepage and J.P. Gazeau, J. Chem.
Phys., 76, 445 (1982).
[26] H.J. Monkhorst and B. Jeriorski,
(1979).
[27] V. Aquilanti and J. Avery,
Int. J. Quantum
71
Int. J.
145
146
Chapter 5
Sturmian Orbitals Applied to
Molecules: an Introduction
Summary
148
5.1 Introduction
Being the Schrodinger equation (SE) for the non-relativistic hydrogen atom analytically soluble in a closed form, the use of hydrogenoid orbitals as basis sets to approximate the wave functions
of atoms and molecules has been extensively considered since the
early days of quantum mechanics. Slater-type orbitals (STO) and
other variants have been proposed and amply tested as alternatives. In order to tackle the obstacles encountered for the ecient
computation of matrix elements (specically the bielectronic integrals), preference has been given to the Gaussian orbitals (GTO),
particularly convenient for such purposes.
149
linear combinations of Slater-type orbitals. The use of the latter
has been superseded by the GTO, for which the slower convergence is often amply balanced by the much simpler algorithm for
the calculation of integrals. Apart from this, the employed techniques are identical. The approach involving Sturmian orbitals will
be shown to proceed along a route fully alternative to modern ab
initio quantum chemistry: it will be seen that there is no need of
iterative Hartree-Fock calculations, because the introduction of the
Sturmian expansion of the electronic wave function leads after algebraic manipulations to a generalized eigenvalue problem (GEP)
where the matrix elements contain mono- and bi-electronic integrals
among Sturmian orbitals and the eigenvalue is not the electronic
energy E but the momentum wave number p0 such that E = p20 =2
(atomic units are implied). The typical integrals of quantum chemistry depend explicitly on the nuclear coordinates Ri , while in the
Sturmian approach parameters si = p0 Ri appear. The basis feature
is therefore that the Born-Oppenheimer separation is not settled before the calculations. The si parameters are instead xed and only
after the solution of the GEP from the relationships Ri = si =p0 one
nds out what is the point in the potential energy surface which
corresponds to an energy p20 =2. Another unusual property of this
approach is that the eigenvalues obtained by solving the GEP come
out as correct solutions not only for ground states, but also for the
excited states corresponding to the same symmetry.
We consider the application of the Sturmian orbital procedure
150
to the ab initio quantum chemistry of molecules as many-electron
many-nuclear systems. The recent proposal, as reviewed in Avery's
book [10, 32], involves a MO-LCAO approach where one-electron
molecular orbitals are computed rst, and then their products are
employed to build-up N -electron congurations. Such an approach
is currently being implemented, one diculty being the recovery
of the Born-Oppenheimer separation, because, as indicated above,
the matching of si parameters and Ri coordinates occurs through
scaling by the eigenvalues p0 .
This chapter presents an account of a promising method for the
application of the Sturmian orbital approach to molecular structure calculations. The new proposal (Sec.5.2.1) for the extension of
generalized Sturmian sets to molecular structure is founded on concepts and formalisms previously elaborated. In the framework of
quantum chemistry jargon, it will be recognized as a valence-bond
type of approach: the calculation of the potential energy curve for
the H2 molecule provides encouraging results already at the level
of minimal basis sets. Sec.5.2.2 is devoted to a discussion of the
practical obstacles to the implementation of this approach, the calculation of integrals, which can be in part circumvented through
the angular momentum algebra applicable in the momentum space
perspective, and in part tackled by exploiting the connections with
STO and the existence of programs dealing with integrals involving
the latter. Conclusion are in Sec.5.3.
151
jxi Xaj
jxj Xp;j j
152
This means that actually there is not only one equation dening the
basis set, because there are many choices for Vpz (x): the basis set
can be divided into several subsets, each of them associated to an
equation like (5.2) having a dierent Vpz(x). A function (x) is a
determinant of Sturmian orbitals u (x Xa ) like Eq.(3.5), provided
that two subsidiary conditions are satised: the rst one is Eq.(3.8),
and the second is a modication of Eq.(3.9) taking into account
that in general Sturmian orbitals appearing in the determinants
are centered on the dierent nuclei of the molecule:
(5.4)
=
p0
+
(5.5)
(5.6)
153
In this case the basis set is constituted by two bielectronic functions;
1 (x) = 1sA(1)1sB (2) corresponds to the potential
V1z (x) =
(5.7)
V2z (x) =
(5.8)
k r
(5.10)
(5.11)
1 2
1 p20
z
z
r + V1 (x) + V2 (x) + r + 2 1sA(1)1sB (2)d1d2 = 0
2
12
154
The latter expression can be simplied with the aid of Eq.(5.2):
Z
(5.12)
1
(1 1 )V1z (x) + V2z(x) +
1sA (1)1sB (2)d1d2 = 0
r
12
Taking into account that the basis set functions associated to the
same operator Vpz (x) enjoy potential weighted orthonormality relationships
Z
2E
dx 0 (x)Vpz(x) (x) = 0 ;
(5.13)
we can modify Eq.(5.12) as follows:
Z
1
2E
z
2E + 1sA(1)1sB (2) V2 (x) +
1s (1)1sB (2)d1d2 (5.14)
1
r12 A
+ 1sB (1)1sA(2) (1
1
z
z
1 )V1 (x) + V2 (x) +
1sA (1)1sB (2)dx = 0
r12
If we write down all the integrals explicitly and exploit Eq.(5.5),
we can rewrite Eq.(5.14) as:
2p0 +p0
1s2B (2)d2
1s2A (1)
jx1 XB j
d1
1s2A(1)d1
+(1
"
"
jx2 XAj d2
pp0 )
2
1s2B (2)
jx1 XAj
jx2 XB j d2
jx1 XB j d1
(5.15)
155
Z
1sA(2)1sB (2)
jx2 XAj d2
"
p0 =
2+
100 (s) + 1
2W100
p0
h
i
p
1 Z 1s2A (1)1s2B (2)
100 (s) = 1 + M100 (s)S 100 (s)
d1d2 2 2M100
(
s
)
S
100
100
100
100
p0
r12
The M's are two-center overlap integrals between Sturmians:
Z
0 l 0 m0
Mnnlm
un0l0m0 (x)unlm(x)dx
For example,
1
M100
100 = 3 e
s (3 + 3s + s2 )
(5.17)
(5.18)
We can show (see Appendix) that these M integrals can be evaluated in momentum space exploiting angular momentum algebra,
analogously to the case of the S integrals encountered in Chapter
4.
Although p0 is in the denition of the exponent k , mono- and
bi-electronic integrals in Eq.(5.16) actually do not depend on the
value of p0 : rather, they depend on the product s = k R (R
being the internuclear distance). Therefore, we pick up a value for
s, compute the matrix elements and then we can get the eigenvalue
156
157
-1
V(R) (a.u.)
-1.05
(c)
-1.1
(b)
-1.15
(a)
3
R (a.u.)
ab initio potential
energy curve for H2 denoted as (a), and the Sturmian approach. Curve
(c) corresponds to the single conguration function in Eq.(5.6) and curve
(b) has been obtained including in the expansion all congurations from
all Sturmian orbitals up to 3d. In the latter case, the variance in the
energy minimum with the one given in [28] is +0.3%.
158
V(R) (a.u.)
-1.1245388
-1.1649344
-1.1723462
-1.174452
-1.1744748
-1.174475
-1.1728541
-1.1685821
-1.1550675
-1.1381316
-1.1201304
-1.1024201
-1.0857874
-1.0706776
-1.0573174
-1.0457865
159
V(R) (a.u.)
-1.1423625
-1.1611956
-1.1694278
-1.1701995
-1.1657679
-1.1577794
-1.1474508
-1.1356935
-1.1231972
-1.1104875
-1.0979637
-1.0859263
-1.0745945
-1.0641206
-1.0545992
-1.046076
-1.0385553
-1.0320072
-1.0263758
-1.0215866
160
161
the method, we explored the possibility of making recourse to the
relationship between Sturmian orbitals and STO.
In practice, since Rico and collaborators have proposed, implemented and made available an ecient program for the calculation
of two-center monoelectronic integrals involving STO, based on the
recurrence relationships which they enjoy [29, 30], and since the
Sturmian orbitals can be given as simple linear combinations of the
STO (explicitly, indicating the latter as nlm :
u1;0;0
u2;0;0
u2;1;m
u3;0;0
u3;1;m
u3;2;m
u4;0;0
u4;1;m
u4;2;m
= 1;0;0 p
= 1;0;0
3 2;0;0
= 2;1;m p
p
= q
2;0;0 + 10 3;0;0
1;0;0 2 3 p
= 2 23 2;1;m
5 3;1;m
(5.19)
= 3;2;m p
p
p
= 1;0;0 3 3 q2;0;0 + 3 10 3;0;0
35 4;0;0
p
p
3
= 5 2;1;m p
5 2 3;1;m + 21 4;1;m
3
= p2 3;2;m
7 4;2;m
dx unlm (x)
jx Xa j un0l0m0 (x Xa0 )
(5.20)
162
The most demanding eort for our valence-bond Sturmian approach appears, as usual in quantum chemistry, when the calculation of the bielectronic multicenter integrals is needed:
Z Z
0
0
jx x0 j u (x )u (x )
C
(5.21)
Our test cases presented in Sec.5.2.1 were tackled in a rather tortuous manner, making recourse to another program due to Rico
and coworkers [31] which in turn in order to obtain integrals for
STO expands them as linear combinations of GTO. The resulting
Gaussian integrals are nally calculated using available techniques.
The ensuing disadvantage of this inelegant procedure is that, since
a satisfactory t of an STO requires about 10 Gaussians, an integral involving four STO's requires the calculation of 104 integrals
among GTO. Moreover our integrals such as (5.21) are linear combinations of integrals among STO: for example an integral involving
four 3s-Sturmians requires a sum over 34 integrals of the STO program, and therefore 34 104 integrals over Gaussians. Therefore
this program, which recommends itself for the accuracy which can
be adjusted due to
exibility in the choice of the number Gaussians required to approximate a STO and satisfactorily covers the
range of permitted quantum numbers (nmax = 6; lmax = 3), although appropriate for test calculations, must be superseded by
more direct approaches to tackle integrals such as Eq.(5.21). Alternatives should also be explored, involving for example elliptic or
other suitable coordinates.
163
164
mentations to systems of increasing complexity. Our current eort
on this problem shows the importance of using momentum space
techniques, which allow the exploit of powerful angular and hyperangular momentum algebra.
ua;n0l0m0 (x)ub;nlm(x)dx
(5.22)
ipR
nlm (
p) n0l0m0 (p)dp
(5.23)
dp =
2k
(5.24)
4 k 5= 2
d
2 2 2 Ynlm (
)Yn0 l0 m0 (
)Yn00 l00 m00 (
)
(p + k )
(5.25)
165
One can easily transform the fourth relationship in Eq.(2.3) into
(1 + cos )2 =
4k5=2 3=2
k
(k2 + p2 )2
(5.26)
b;n0 l0 m0
a;nlm
M
Z
(5.27)
d
(1 + 2 cos + cos2 )Ynlm (
)Yn0 l0 m0 (
)Yn00 l00 m00 (
)
Ynlm (
)Yn0 l0 m0 (
) =
N;L;M
"
#
nn0 N (2l + 1)(2l0 + 1) 1=2
2 2
8 0
n
>
>
< 2
1
0
n 1
2
l0
n 1
2
n 1
(5.28)
N 1 >
=
2 >
N 1
Y
NLM
>
>
;
(
)
l
L
From a three-term recurrence relationship involving Gegenbauer
polynomials [38], we have obtained the following expression:
:
cos YNLM (
) =
v
u
u
t
1 (N L)(N + L + 1)
YN +1;LM (
) (5.29)
2
N (N + 1)
v
u
L 1)
1u
t (N + L)(N
YN 1;LM (
)
2
N (N 1)
Inserting Eqs.(5.28) and (5.29) into Eq.(5.27), we get a solution
for the integral:
b;n0 l0 m0
a;nlm
!
"
#
2 3=2 X nn0 N (2l + 1)(2l0 + 1) 1=2
=
k
2 2
NLM
(5.30)
166
8 0
> n
>
< 02
("
1+
l0
n 1
2
n 1
N 1 >
=
2 >
N 1
>
>
;
L)(N + L + 1) (N + L)(N L 1)
+
(1 N;1 )
4N (N + 1)
4N (N 1)
(N
v
u
u
t
uNLM (x)
v
u
u
t
v
u
4(N
(N
L)(N + L + 1)
uN +1;LM (x)
N (N + 1)
(N + L)(N L
N (N 1)
1 u
t (N
+
4(N + 1)
+
1
n 1
2
1)
v
u
u
t
1)
L2 ]
(N + L)(N
1)2
L 2)[(N
N (N 2)
uN +2;LM (x)
L2 ]
Bibliography
[1] H. Shull and P.-O. Lowdin,
[2] V. Fock,
J. Chem. Phys.
, 98,145 (1935).
Z. Phys.
J. Chem. Phys.
167
[7] V. Aquilanti, , S. Cavalli, C. Coletti, D. Di Domenico, and
G. Grossi. Quantum Systems in Chemistry and Physics. Vol I:
Basic Problems and Model Systems, page 289 (2000).
[8] J. Avery, Hyperspherical Harmonics, Applications in Quantum
Theory. Kluwer Academic Publishers, Dordrecht, The Netherlands
(1989).
[9] T. Shibuya and C.E. Wulfman, Proc. Roy. Soc., A286, 376 (1965).
[10] J. Avery, Hyperspherical Harmonics and Generalized Sturmians.
Kluwer Academic Publishers, Dordrecht, The Netherlands (2000).
[11] J. Avery,
, 214, 1
Chem. Phys.
J. Chem. Phys.
Opt. Spectrosc.
Chem. Phys.
J. Phys. A
J. Chem.
[19] H.J. Monkhorst and B. Jeriorski, J. Chem. Phys., 71, 5268 (1979).
[20] V. Aquilanti and J. Avery,
, 267, 1 (1997).
,
168
[22] V. Aquilanti, A. Caligiana, S. Cavalli and C. Coletti, Int. J. Quantum Chem., 92, 212, (2003).
[23] P. M. Morse and H. Feshbach,
McGraw-Hill, New York (1953).
J. Chem. Phys.
J. Chem. Phys.
, 39, 71 (2001).
J. Chem. Phys.
J. Comp. Chem.
Comp. Phys.
[34] R. Szmytkowski,
J. Phys. B
[35] R. Szmytkowski,
J. Phys. A
[36] R. Szmytkowski,
Phys. Rev. A
[37] R. Szmytkowski,
J. Phys. B
and
(2002).
[38] M. Abramowitz and I.A. Stegun. Handbook of Mathematical Function. Dover, New York, 1964.
Chapter 6
Anharmonic frequencies and Berry
pseudorotation motion in PF5
Summary
We study the dynamics of PF5 , using a density functional potential energy surface. The full surface has twenty D3h minima
connected by thirty C4v transition states. We use the theory of the
molecular symmetry group to understand the tunnelling splittings
of the vibrational levels. We estimate a value for the splitting by
studying the motion between two equivalent minima. This splitting is very small for the ground vibrational state. Therefore we
then study the vibrations of PF5 in one of its minima, using normal
coordinates with anharmonic perturbation theory and variational
theory. Our predicted fundamental frequencies agree well with observation. Finally we use the theory of the Reaction Path Hamiltonian to study the interaction of the vibrational modes with a large
169
170
amplitude motion connecting two minima.
6.1 Introduction
In this chapter, we shall study the Berry pseudorotation [1] dynamics of PF5 . Phosphorus penta
ouride has twenty equivalent
D3h minima which are connected by thirty C4v transition states.
Firstly we shall look at the structure of the vibrational energy levels including the eects of the multiple tunnellings, using the theory of the molecular symmetry group. We shall estimate a value
for the splittings which turns out to be very small for the ground
state. Then we shall look at the dynamics in the region of one
minimum, generating a quartic normal coordinate potential energy
surface using density functional theory. We use both perturbation
theory and variational theory to predict fundamentals, overtones
and combination bands.
We draw attention to some earlier studies on PF5 . Tschumper,
Fermann and Schaefer report structures, thermochemistry and electron anity using a variety of density functionals [2]. Wasada and
Hirao have studied the minimum and the transition state of the
pseudorotation using ab initio quantum chemistry [3]. They report
a barrier which is estimated between 1480 cm 1 and 1770 cm 1 .
Breidung and Thiel have used SCF theory to calculate an anharmonic force eld for PF5 . This reference provides a careful analysis
of the existing experimental data. The level of agreement they
171
obtained for the fundamentals (with MP2 for the quadratic force
constants) is similar to the accuracy we shall obtain using density
functional theory.
172
l eq
l eq
l ax
l pi
l ot
l ax
l eq
l ax
l ot
l eq
l ot
2
x
D 3h
l eq
4
4
y
l eq
l ot
l ax
3
C 4v
D 3h
173
Table 6.1: DFT optimization of PF5 ground state and the Berry
pseudorotation saddle point. (a Ref. [5]).
Expt.a
1.580
1.535
1.575
1.531
1.535
1.579
102.4
1.535
1.580
102.4
1.531
1.575
102.4
1279
1292
1271
1.577 0.005
1.534 0.004
174
Table 6.2: Calculated harmonic frequencies (cm 1 ) at the D3h minimum, compared with observed fundamentals from a Ref. [22] and
references therein.
Mode (D3h ) B97-1 B3LYP B3P91 Expt.a
1 (A01 )
792
783
792
816
0
636
626
634
648
2 (A1 )
00
3 (A2 )
934
920
933
946.4
00
4 (A2 )
559
557
560
575
0
5 (E )
1001
988
1000 1025.7
0
511
509
511
533
6 (E )
0
7 (E )
157
156
157
174
00
496
493
496
512
8 (E )
175
Table 6.3: Calculated harmonic frequencies in cm 1 for the C4v
transition state.
Mode (C4v ) B97-1 B3LYP B3P91
1 (A1 )
993
981
994
784
775
783
2 (A1 )
501
498
503
3 (A1 )
4 (B1 )
630
618
628
5 (B1 )
113i
117i
114i
6 (B2 )
558
554
558
966
949
964
7 (E )
8 (E )
552
547
552
356
354
357
9 (E )
and one of the three equatorial bonds (we call it `pivot') will be
almost unaected by the BP. Since there are three possible choices
for the pivot, each minimum (D3h ) is connected to the three nearest
equivalent minima on the surface by three equivalent paths. Thus
we are dealing with a potential energy surface having 20 equivalent
D3h minima and (20 3)=2 = 30 equivalent C4v transition states
(TS) connecting them in pairs. In Fig. 3 we report the Balaban
graph corresponding to our surface [10]. In the literature there
are several isomorphic versions of this graph, and we have tried to
choose an easily readable one.
We can determine the splitting patterns, i.e. the number and
the degeneracy of the splitting levels, by group theoretical considerations. If we neglected the presence of multiple minima, we could
176
C3 S 3
v
v
v
v
C2
C2
C2
C2
C4
C2
D3h
trigonal bypiramid
C 2v
intermediate configuration
C 4v
transition state
177
12
5
34
45
4
2
1
35
13
25
23
15
5
14
3
2
24
4
24
14
5
15
23
25
13
35
1
2
45
34
3
12
molecule XY5 . The D3h isomers are labeled by the numbers of the two
axial ligands, putting rst the one from which the equatorial ligands are
seen in clockwise sequence. If the two numbers are in decreasing order,
we invert and oversign them; such an isomer is the enantiomer of the not
oversigned one. The pivot which allows the passage between two isomers
has the number that does not appear in the labels of the isomers.
178
label the vibrational wavefunctions according to the irreducible representations of the D3h group. But when we want to consider the
interactions among several equivalent structures, we have to exploit
the molecular symmetry (MS) group to predict the degeneracy of
splitting levels and to label them [11]. The complete nuclear permutation inversion (CNPI) group of PF5 has 240 elements, being
the product of the complete permutation group (5! elements) and
the inversion group (2 elements). The MS group is a subgroup of
the CNPI group, obtained from the latter after deleting all the unfeasible elements, i.e. all the elements that interconvert numbered
equilibrium forms of the molecule separated by an insuperable barrier in the potential energy surface. For our molecule there is no
unfeasible element in the CNPI group, so this coincides with the MS
group. Then in order to predict the splitting patterns we have to
obtain the correlation between the A01 irreducible representation of
the D3h group and the corresponding irreducible representations in
the CNPI group G240 of PF5 . This was achieved using a code written by D. J. Wales [12]. The result is that the twenty ground state
vibrational levels span 6 irreducible representations: 2 of them are
non-degenerate, 2 are 4-fold degenerate and 2 are 5-fold degenerate.
In order to provide a quantitative estimation of the tunnel splittings, we exploit an approximate method described by Wales [12,
13]. This approach is an analogue to the LCAO method for the
construction of molecular orbitals. If there are n distinguishable
versions of a given structure, then there are n degenerate wave
179
functions, each localized on one particular potential energy well,
for any given vibrational state of the molecule. When rearrangements occur only between these n distinguishable versions, we may
approximate the true wave function as a linear combination of localized wave functions. We then construct the secular equations. In
the case of PF5 , there is just one type of degenerate rearrangement
which links 20 versions of the same structure. All the diagonal elements of the Hamiltonian H^ are the same (), and the o-diagonal
elements of H^ are denoted by , the tunneling matrix element, if
the corresponding minima are directly linked by the rearrangement
(adjacent), and are set zero otherwise. Then the splitting levels
are the eigenvalues of a 20x20 matrix which in each row has 3 odiagonal elements equal to and 16 equal to zero. Diagonalization
yields the splitting pattern for the ground state of PF5 , and the result, represented in Fig. 4, is seen to be consistent with the group
theoretical predictions.
Next we have to provide a numerical (approximated) value for
. Let us consider the equation
!
1 d2
+ V (Q) Ei 'i = 0
2 dQ2
(6.1)
180
(1)
+3
(4)
+2
A1
(5)
D3h
(5)
(4)
(1)
(20)
240
Figure 6.4: The splitting pattern for the vibrational ground state of
PF5 , taking into account the multiple equivalent minima. The numbers
between parentheses indicate the degeneracy of levels.
181
and 5 to be on the yz plane, and atoms 2 and 4 to be on the xy
plane (referring to Fig. 1); c) x a value for ; d) optimize all the
other internal degrees of freedom (we use the functional B3P91 and
basis set TZ2P). Eq.(6.1) is solved numerically, and the dierence
between the two lowest-lying levels is used to evaluate (Fig. 5):
E
' 2
E1
= 0:06 cm 1
(6.2)
Our estimated value of is indeed very small, and in fact no splitting due to Berry pseudorotation has been detected experimentally. Takami and Kuze have studied the 3 fundamental and the
3 + 7 7 hot band using infrared absorption spectroscopy. They
state that they do not observe any splitting attributable to BP with
their 10 MHz resolution [14]. Of course our calculation gives only
a crudest estimate for the splitting.
However, we nd that there will be a substantial splitting for
the higher vibrational levels. Table 2 shows that the magnitude of
the frequency of vibration which corresponds to interconversion is
160 cm 1 , i.e. seven quanta, suce to reach the top of the barrier. Therefore if it were possible to create such a highly excited
vibration, then the eects of BP would be observable.
182
Energy (Eh)
-840.72
-840.73
C4v
E2
-840.74
D3h
-400
E1
-200
D3h
0
200
Q (a.u.)
Figure 6.5: A C2v reaction path connecting a D3h minimum and a C4v
400
183
in isolation, using the B3P91 functional and the TZ2P basis set
to determine the anharmonic force eld. We expand the potential
energy surface through quartic terms:
1 X
1X 2 1X
!i qi +
ijk qi qj qk +
qqq q
(6.3)
2 i
6 ijk
24 ijkl ijkl i j k l
In this standard notation !i are the harmonic wavenumbers, qi
are the dimensionless normal coordinates, and ijk , ijkl are the
third and fourth order force constants, respectively. If Eq. (6.3)
corresponds to a Taylor series, then we use this identication:
V =
@ 3 V
ijk =
@qi @qj @qk 0
(6.4)
@4V
ijkl =
(6.5)
@qi @qj @qk @ql 0
As usual the third and semi-diagonal fourth order (iijk ) force
constants were evaluated by nite dierence calculations from analytical second derivatives obtained at the equilibrium geometry
and at the geometries displaced from equilibrium along the normal
modes:
(6.6)
(6.7)
ijk =
iijk =
184
We have discussed elsewhere how the magnitude of the displacements are chosen, and subsequent averaging [15]. Essentially the
displacement was chosen to be 100 Eh over the equilibrium geometry.
To obtain the vibrations on this potential energy surface, we
may use either perturbation theory or variational theory. Secondorder perturbation calculations were carried out using the program
`SPECTRO' [16]. In a simplest form, SPECTRO obtains the energy levels of an asymmetric top i.e. it determines the anharmonic
constants xkl
E (v) =
1
1
1 X
!k (vk + ) + xkl (vk + )(vl + );
2 kl
2
2
(6.8)
185
Table 6.4: Fundamental frequencies (cm 1 ) from a quartic force
eld (DFT B3P91/TZ2P, Eq. 6.3), calculated by b perturbation
theory (SPECTRO) and cvariational theory (MULTIMODE). a See
Ref. [22] and references therein.
Mode (D3h )
1 (A01 )
2 (A01 )
3 (A002 )
4 (A002 )
5 (E 0 )
6 (E 0 )
7 (E 0 )
8 (E 00 )
PTb
786
629
926
557
991
512
157
493
VTc a Expt.
787 816
630 648
927 946.4
557 575
992 1025.7
512 533
157 174
493 512
186
(SCF) calculations, which were followed by conguration interaction (CI) using the orthonormal virtual expansion functions of the
SCF zero-point level. The `MULTIMODE' package has been described previously [19]. `MULTIMODE' uses numerical integration
for the matrix elements. We use the parameter Icoupl = 3 which
means that the code properly considers all terms in our potential
V . The Coriolis terms are evaluated up to the case when 3 normal
coordinates are simultaneously nonzero.
The SCF vibrational states are formed from a basis of NVF=6
optimised one-dimensional functions in the 12 normal modes. These
are in turn obtained from an initial set of 12 primitive harmonic oscillator basis functions, integrated over 15 points by Gauss-Hermite
quadrature, in matrix elements constructed from one-dimensional
cuts of the potential. For each required state the basis is diagonalised for each mode in turn in the eld of the remaining modes
in their respective states until self-consistency is reached.
For improved descriptions of the vibrational states we perform
CI. The CI basis is itself built from 1-mode, 2-mode, etc. expansion
sets in all normal modes of the molecule, i.e. from sets allowing
simultaneous excitation of one, two etc. modes. Each category is
subject to a maximum quantum Q for each mode and a maximum
sum S over all modes. We report calculations using four modes,
with Q=6 and S=6, in Table 6.4.
Table 4 shows a remarkable agreement between the perturbation
theory and the variation theory predictions for the fundamentals.
187
This is hardly surprising; when using the data from Table 2, it is
seen that the eects of anharmonicity are very small (less than 10
wavenumbers for each mode). The anharmonic eects are smaller
than the errors arising from the use of DFT, but even those do not
exceed 35 wavenumbers. Therefore the fundamentals of PF5 are
well reproduced by considering it as a rigid isolated molecule.
188
determination of the `Reaction Path' connecting two adjacent minima (using DFT). Ideally we should nd the steepest descent path
from the TS to the minima (it is symmetric about the TS); for
simplicity we determined a closely related path which is parameterised by the angle (see g 1), enforcing C2v symmetry. The DFT
energy is minimised with respect to all other internal coordinates.
This gives the same reaction path as obtained in section 3. At each
point on the path (which is symmetric about the transition state),
the projected force constant matrix is determined, giving the harmonic frequencies and the associated vectors, which are required for
the RPH implementation. Our representation of the reaction path
through the angle allows an automatic continuation of it beyond
the minima by further reducing the angle . The greatest diculty
with the application of RPH is that the vibrational modes must
continuously change along the path, in particular at an accidental
crossing. This was carefully checked, and we also ensured that all
frequencies and vectors were symmetrical across the TS.
Eectively this entire calculation is an attempt to model the role
of the multiple minima on the vibrational dynamics. From section
3 we know the eects on vibrations along the paths connecting the
multiple minima, with levels at E0 ; E0 2; E0 3 . Here we
shall only obtain E0 . We therefore expect the zero point energy
to be split in this way, and indeed we expect all the vibrational
levels to be split, i.e. the number of levels will be doubled. Because
the reaction path motion is degenerate at the minima, we expect
189
the this motion to be split into 4, with two near pairs, corresponding
to motion along the path and motion perpendicular to the path. If
we were able to study the full problem then each original level E0
would be split into 6 levels, according to the theory of section 3.
We believe that the magnitude of the splittings should be in the
ratios given by section 3.
The results of the RPH calculations are given in Table 6.5.
The original MULTIMODE calculations are also given (as VT)
for comparison, in order to demonstrate the splittings discussed
above. The VT calculations use only the full quadratic force eld
at the minima (i.e. no anharmonicity); the RPH calculations use
only a quadratic force eld perpendicular to the path (i.e. only
anharmonicity along the path). The table shows the predicted levels for the zero point, the fundamentals, the overtone 27 and the
combination bands 7 + k (k=1-6,8). 7 corresponds to the (degenerate) motions at the minima, one of which corresponds to the
reaction path. All the expected splittings are present. Highlights
are
(i) the zero point motion is split by 0.13 cm 1 .
(ii) the degenerate mode 7 is split into two motions: (162.17,
162.54), which is perpendicular to the path (this is a fundamental vibration), and (195.75,197.40), which corresponds to the rst
excited vibration along the path, with a consequent larger splitting.
(iii) a nondegenerate fundamental, such as 1 is split (798.79,798.92)
by 0.13 cm 1 .
190
(iv) the combination 8 + 7 , both originally degenerate, now
correspond to 8 levels, easily identiable as 4 related pairs.
We therefore believe that these RPH calculations have gone
some way towards providing a numerical demonstration of the expected vibrational levels for PF5 .
6.6 Conclusion
In this chapter we have studied the dynamics corresponding to the
low lying vibrations of PF5 . We have used density functional theory
to determine the important aspects of the potential energy surface,
a theory which we believe is entirely adequate for these purposes,
bearing in mind that we do not study parts of the surface which
involve the breaking of bonds. We have calculated the barrier to
interconversion between equivalent minima to be 1280 cm 1 .
Firstly we studied the eects of tunnelling motion on the vibrations around the twenty minima which are connected by thirty
transition states. We used the Molecular Symmetry group to predict the degeneracies of the vibrations; we also used a Huckel type
analysis for this purpose. We predicted that the splittings of these
vibrations are very small.
Secondly we studied the anharmonic motion of PF5 around one
of its minima, using both perturbation theory and variational theory. We compared our predictions for the fundamentals with those
available in the literature, agreement to 15 cm 1 being obtained.
191
Finally we use the Reaction Path Hamiltonian to study the vibrations with one large amplitude motion connecting two adjacent
minima. We were able to interpret the resultant splittings of the vibrations and would also predict the nature of the vibrational levels
if it were possible to include the eect of all the minima simultaneously.
To sum up, the principal weakness of the work presented in this
chapter is the use of Density Functional Theory to generate the
potential energy surfaces. We were interested in a study of the dynamics, and methods for this, and we believe that our methods are
completely appropriate to a more accurate potential energy surface.
From the calculations of Wasada and Hirao [3] one possibility is that
our transition state is too low. However a much greater criticism
is that we have only been able to calculate the interactions of one
minima with another, and not the interactions of one with minima
with three other minima. The eect of including all the interactions simultaneously will be to substantially reduce the tunnelling
splitting, but at present we have no numerical way of determining
this, which should require that the large amplitude motion be occurring on at least a two-dimensional surface, to accomplish for the
trifurcation of the reaction path from each minimum towards its
neighbours.
192
Bibliography
[1] R. S. Berry,
J. Chem. Phys.
J. Chem.
[4] R.D. Amos, I.L. Alberts, J.S. Andrews, S.M. Colwell, N.C. Handy,
D. Jayatilaka, P.J. Knowles, R. Kobayashi, G.J. Laming, A.M. Lee,
P.E. Maslen, C.W. Murray, P. Palmieri, J.E. Rice, E.D. Simandiras,
A.J. Stone, M.-D. Su, and D.J. Tozer, Cadpac6.5, The Cambridge
Analytic Derivatives Package (University of Cambridge, UK), 2000.
[5] L. S. Bartell, K. W. Hansen,
, 4, 1775, 1965.
Inorg. Chem.
J. Chem. Phys.
J. Chem. Phys.
, 66, 1, 1977.
[12] D. J. Wales,
[13] D. J. Wales,
, Academic
J. Chem. Phys.
193
[16] J.F. Gaw, A. Willetts, W.H. Green, and N.C. Handy, in "Advances
in Molecular Vibrations and Collision Dynamics", ed. J.M. Bowman, JAI Press, Greenwich, CT (1990).
[17] D. Papousek, M.R. Aliev, Molecular Vibrational-Rotational Spectra, Academia, Prague (1982).
[18] J. K. G. Watson,
Mol. Phys.
[20] D. P. Tew, N. C. Handy and S. Carter Mol. Phys., 99, 393 (2001).
[21] D. P. Tew, N. C. Handy and S. Carter J. Chem. Phys., submitted
[22] J. Breidung, W. Thiel,
J. Mol. Struct.
194
Table 6.5: Fundamental frequencies, and overtones and combination bands involving 7 calculated with Multimode (VT) and the
reaction path Hamiltonian (RPH). Only harmonic terms are included in the normal coordinate forceeld
Mode (D3h )
VT
RPH
0
1 (A1 )
792
798.79,798.92
0
635
640.01,640.12
2 (A1 )
3 (A002 )
936
938.95,939.05
00
4 (A2 )
562
565.98,566.16
0
5 (E )
1001;1001
1004.15,1004.33;1006.40,1006.59
0
6 ( E )
515;516
505.54, 505.76;519.31,519.44
0
162;163
162.17,162.54;195.75,197.40
7 (E )
00
8 (E )
498;499
500.78,500.86;502.61,502.77
27
324;325;325
321.24,324.27;366.99,368.98;384.57,388.70
943
961.23,962.64;992.68,993.23
1 + 7
2 + 7
797;798
802.45,803.82;835.51,836.10
3 + 7
1098;1098
1101.37,1102.78;1135.71,1136.29
4 + 7
724;725
729.95,730.75;764.76,766.07
5 + 7
1163;1164;1164;1164
1167.43,1168.74;1199.76,1200.33
6 + 7
678;678;678;679
667.63,669.13;681.47,683.33;703.07,703.79;718.18,718.92
660;661;661;661
662.77,664.57;666.77,667.26;694.16,694.66;703.95,705.04
8 + 7
Contents
195
196
Conclusions
Here we summarize the main results exposed in this work. The
chapters of this thesis originated several publications that are listed
in the Bibliography of the present section.
In chapter 1 [1] we have studied the four dierent types of Sturmian orbitals arising from the separation of variables in Schrodinger
equation for the hydrogen atom: polar, parabolic, spheroidal and
spheroelliptic. The denition of spheroidal and spheroelliptic Sturmians includes a continuous parameter, which allow them to perform smooth transitions respectively from polar to parabolic Sturmians and from polar Sturmian oriented towards an axis to polar
Sturmian oriented towards a dierent axis. The symmetry properties of the alternative basis sets and the orthonormal transformations to each other have been investigated.
In chapter 2 [2] we have studied the corresponding problem for
Sturmians in momentum space (there are six types of them: spherical, Stark, Zeeman, spheroelliptic, elliptic cylindrical of type I and
II), exploiting in our investigation the method of trees and the
close relationship with hyperspherical harmonics. Then, we have
197
198
discussed the connections among basis sets in the conguration
and momentum space. Spherical, Stark and spheroelliptic Sturmians in momentum space are respectively the counterparts of polar,
parabolic and spheroelliptic Sturmians in conguration space. Elliptic cylindrical Sturmians of type I and II include a continuous
parameter that allows them smooth transitions between Stark and
Zeeman functions, and spherical and Zeeman functions respectively.
In chapter 3 we have applied one type of Sturmian orbitals in
conguration space (the polar ones) to the solution of Schrodinger
equation for atoms. The consequences arising from the use of Sturmians has been exposed: the kinetic energy operator disappears
from the secular equation; the representation of the nuclear attraction potential is diagonal; the eigenvalues are not the energies of the
p
2E ; since the orbitals are not orthogonal, matrix
system, but
elements are to be calculated in the framework of Slater-Condom
generalized rules; the eect of external elds could be easily incorporated in our formalism. A calculation on Helium atom electronic
states has shown that the description of excited states is very good
allowing one to simulate easily the atomic spectrum. This is at
variance with conventional quantum chemistry methods. An application of the method in the case of external eld has been performed
in order to compute the rst order static polarizabilities of Lithium.
Chapter 4 deals with Sturmians and one-electron molecules [3].
Analytical formulas were obtained for nuclear attraction integrals
involving polar and parabolic Sturmians. A calculation of the po-
199
tential energy curve of H+2 has been performed in order to test the
convergence properties of Sturmian basis sets at dierent levels of
iteration of the secular equation.
In chapter 5 [4, 5] we have proposed an original method for
solving the molecular Schrodinger equation for molecules, inserting
Sturmian orbitals into a valence bond approach. We have not been
able to solve analytically multicenter repulsion integrals involving
Sturmians, so we had to expand the latters as linear combinations
of Gaussian functions and solve numerically the integrals. We calculated the potential energy curve for H2 , demonstrating that such
a method can be practically implemented.
Chapter 6 [6] is not closely related to the previous ones. We
studied the dynamics of PF5 , using a density functional potential
energy surface. The full surface has twenty D3h minima connected
by thirty C4v transition states. We used the theory of the molecular
symmetry group to understand the tunnelling splittings of the vibrational levels. We estimated a value for the splitting by studying
the motion between two equivalent minima. This splitting is very
small for the ground vibrational state. Therefore we then studied
the vibrations of PF5 in one of its minima, using normal coordinates
with anharmonic perturbation theory and variational theory. Our
predicted fundamental frequencies agree well with observation. Finally we used the theory of the Reaction Path Hamiltonian to study
the interaction of the vibrational modes with a large amplitude motion connecting two minima.
200
Bibliography
[1] V. Aquilanti, A. Caligiana and S. Cavalli,
Chem., 92, 99, (2003).
Int. J. Quantum
Int. J.