Buscar

(Lecture Notes in Chemistry 74) Dr. Mireille Defranceschi, Prof. Claude Le Bris (auth.) Mathematical Models and Methods for Ab Initio Quantum Chemistry Springer Verlag Berlin Heidelberg (2000)

Prévia do material em texto

Lecture Notes in Chemistry 
Edited by: 
Prof. Dr. Gaston Berthier 
Universite de Paris 
Prof. Dr. Hanns Fischer 
Universitat ZUrich 
Prof. Dr. Kenichi Fukui 
Kyoto University 
Prof. Dr. George G. Hall 
University of Nottingham 
Prof. Dr. JUrgen Hinze 
Universitat Bielefeld 
Prof. Dr. Joshua Jortner 
Tel-Aviv University 
Prof. Dr. Werner Kutzelnigg 
Universitat Bochum 
Prof. Dr. Klaus Ruedenberg 
Iowa State University 
Prof Dr. Jacopo Tomasi 
Universita di Pisa 
74 
Springer-Verlag Berlin Heidelberg GmbH 
M. Defranceschi C. Le Bris 
Mathematical 
Models and Methods 
for Ab Initio 
Quantum Chemistry 
Springer 
Authors 
Dr. Mireille Defranceschi 
CEA-Saclay, Bât. 125 
DPE/SPCP/LEPCA, 
91191 Gif-sur-Yvette Cedex, France 
E-mail: defrancesc@camac.cea.fr 
Prof. Claude Le Bris 
C.E.R.M.I.C.S. 
Ecole Nationale des Ponts et Chaussees 
6&8 Avenue Blaise Pascal 
Cite Descartes, Champs sur Mame 
77455 Mame La VaII ee Cedex 2, France 
E-mail: lebris@cermics.enpc.fr 
Cataloging-in-Publication Data applied for 
Die Deutsche Bibliothek - CIP-Einheitsaufnahme 
Mathematical models and methods for ab initio quantum chemistry I 
Mireille Defranceschi ; Claude LeBris. - Berlin; Heidelberg ; New 
York; Barcelona; Hong Kong; London; Milan; Paris; Singapore; 
Tokyo: Springer, 2000 
(Lecture notes in chemistry ; 74) 
ISBN 978-3-540-67631-7 ISBN 978-3-642-57237-1 (eBook) 
DOI 10.1007/978-3-642-57237-1 
ISSN 0342-4901 
ISBN 978-3-540-67631-7 
This work is subject to copyright. AII rights are reserved, whether the whole or part 
of the material is concemed, specifically the rights of translation, reprinting, re-use 
of illustrations, recitation, broadcasting, reproduction on microfilms or in any other 
way, and storage in data banks. Duplication of this publication or parts thereof is 
permitted only under the provisions of the German Copyright Law of September 9, 
1965, in its current version, and permission for use must always be obtained from 
Springer-Ver1ag. Violations are liable for prosecution under the German Copyright 
Law. 
© Springer-Verlag Berlin Heidelberg 2000 
Originally published by Springer-Verlag Berlin Heidelberg New York in 2000 
Softcover reprint ofthe hardcover Ist edition 2000 
The use of general descriptive names, registered names, trademarks, etc .. in this 
publication does not imply, even in the absence of a specific statement, that such 
names are exempt from the relevant protective laws and regulations and therefore 
free for general use. 
Typesetting: Camera ready by author 
Printed on acid-free paper SPIN: 10771679 51/3142 - 543210 
Contents 
Foreword, by Mireille Defranceschi & Claude Le Bris ix 
I General topics 1 
1 Is a molecule in chemistry explicable as a broken symmetry in 
quantum mechanics?, by B. Sutcliffe 3 
1 Introduction.................... 4 
2 The equations of motion for the molecule . . . . 5 
3 The Eckart choice of a frame fixed in the body . 7 
3.1 The effects of nuclear permutations on the Eckart coordinates 9 
3.2 Discussion and conclusions. . . . . . . . . . . . . . . . . . 14 
2 SCF algorithms for HF electronic calculations, by E. Cance.s 17 
1 Introduction.................... 18 
2 A brief presentation of the Hartree-Fock model. 19 
3 General remarks on SCF algorithms. . . . . . . 24 
4 The Roothaan algorithm: why and how it fails. 27 
5 Level-shifting........... 31 
6 The Optimal Damping Algorithm . . . . . . . . 33 
3 A pedagogical introduction to Quantum Monte-Carlo, 
by M. Caffarel & R. Assam! 
1 Introduction....... 
2 QMC for a 2 x 2 matrix .. . 
3 
2.1 The 2 x 2 matrix .. . 
2.2 QMC is a power method 
2.3 Introducing a stochastic matrix 
2.4 Pure Diffusion Monte Carlo (PDMC) 
2.5 Diffusion Monte Carlo (DMC) .... 
2.6 Stochastic reconfiguration Monte Carlo (SRMC) 
2.7 Accelerating QMC .... 
QMC beyond the 2 x 2 matrix .. 
3.1 QMC for a N x N matrix 
3.2 QMC in a continuum ... 
45 
46 
49 
49 
50 
51 
55 
56 
58 
61 
66 
66 
66 
VI 
4 On the controllability of bilinear quantum systems, by G. Turinici 75 
1 Introduction . . . . . . . . . . . . . 76 
2 Infinite dimensional controllability. . . . 
3 Finite dimensional controllability .... 
4 Transfer graph and necessary conditions 
5 Controllability results. . . . . . . 
5.1 Local exact controllability 
5.2 Global controllability ... 
6 Application to a five level system 
7 Conclusions 
8 Appendix ............ . 
II Condensed phases 
76 
78 
79 
80 
80 
83 
86 
88 
88 
93 
5 Recent mathematical results on the quantum modeling of crystals, 
by 1. Catto, C. Le Bris & P-L. Lions 95 
1 Introduction................ 96 
2 Models from Density Functional Theory 98 
3 Hartree type models ........... . 102 
4 Hartree-Fock type models ........ . 104 
5 The optimal periodic configuration for the nuclei. . 110 
6 Extensions and perspectives . . . . . . . . . . . . . 113 
6 Local density approximations for the energy of a periodic Coulomb 
model, by O. Bokanowski, B. Grebert & N. J. Mauser 121 
1 Introduction................ . 122 
2 Deformations and the Jacobian problem .......... . 123 
3 Kinetic energy estimates . . . . . . . . . . . . . . . . . . . . 124 
4 Justification of the X a method and exchange energy approximation . 126 
5 Global energy bound ........................... 127 
7 A mathematical insight into 
ab initio simulations of the solid phase, by X. Blanc 133 
1 Introduction..... 
· 134 
2 Crystalline structure . 
· 134 
3 Band theory . . . . . . 
· 136 
3.1 Band structure 
· 136 
3.2 Density of states 
· 138 
4 Electronic calculations: standard approximations 
· 140 
4.1 Hartree-Fock formalism ... . 
· 140 
4.2 Density functional theory .. . 
· 142 
4.3 The use of pseudo-potentials . . 
· 143 
4.4 Self-consistent field algorithms. 
· 144 
5 Numerical schemes 
· 145 
5.1 Basis sets . . . . . . . . . . . . 
· 145 
6 
5.2 
5.3 
Reciprocal space integration and special point technics 
Ewald sums 
Conclusion. . . ..................... . 
vii 
147 
· 149 
· 152 
8 Examples of hidden numerical tricks in a solid state determination 
of electronic structure., by M. Defranceschi f3 V. Louis-Achille 159 
1 Introduction...................... . 160 
2 General aspects of the Density Functional Theory (DFT) . 160 
3 Choice of the method ......... 162 
4 
5 
3.1 Representation of the function . 163 
3.2 Use of pseudo-potentials . . 164 
Choice of the number of k-points . 165 
Other numerical tricks · 165 
5.1 Results .......... . · 166 
9 Quantum mechanical models for systems in solution, 
by B. Mennucci 171 
1 Introduction.. ................. . · 172 
2 The Effective Hamiltonian and the Free Energy · 172 
3 Intermolecular Forces . . . . . . . · 175 
3.1 Repulsion and Dispersion ....... . · 176 
3.2 Electrostatic contribution ........ . 178 
4 Quantum Mechanical Theory of Solvation: the ASC approach . 185 
5 The solute cavity and its tessellation . 187 
6 Derivatives and Molecular Properties . 189 
6.1 Free Energy Surfaces . 189 
6.2 External Properties. . . . . . 192 
7 An example of applications. . . . . . . 195 
7.1 Structures and (hyper)polarizabilities of push-pull systems in 
solution . 196 
8 Conclusions................................ 202 
III Relativistic models 209 
10 Variational methods in relativistic quantum mechanics: new ap-
proach to the computation of Dirac eigenvalues, by J. Dolbeault, 
M. J. Esteban f3 E. Sere 211 
1 Introduction............... . . . . . . 212 
2 Min-max approaches. . . . . . . . . . . . 214 
3 
4 
Minimization method and corresponding min-max approaches. 
3.1 A constrained minimization method. . 
3.2 Relationship with Talman's min-max .. 
Some related Ilumerical computations. 
· 216 
· 216 
.217· 218 
Vlll 
11 Quaternion symmetry of the Dirac equation, by 
H. J. Aa. Jensen 
1 Introduction . . . . . . . . 
2 
3 
4 
5 
The Dirac equation . . . . 
The full symmetry group . 
3.1 Spatial symmetry. 
3.2 Time reversal symmetry 
The quat ern ion Dirac equation. 
Discussion and conclusion . . . 
T. Saue fj 
227 
· 228 
.228 
· 231 
· 231 
· 236 
.240 
· 243 
Foreword 
On the occasion of the fourth International Conference on Industrial and Applied 
Mathematics!, we decided to organize a sequence of 4 minisymposia devoted to the 
mathematical aspects and the numerical aspects of Quantum Chemistry. Our goal 
was to bring together scientists from different communities, namely mathematicians, 
experts at numerical analysis and computer science, chemists, just to see whether 
this heterogeneous set of lecturers can produce a rather homogeneous presentation 
of the domain to an uninitiated audience. 
To the best of our knowledgde, nothing of this kind had never been tempted 
so far. It seemed to us that it was the good time for doing it, both .because the 
interest of applied mathematicians into the world of computational chemistry has 
exponentially increased in the past few years, and because the community of chemists 
feels more and more concerned with the numerical issues. 
Indeed, in the early years of Quantum Chemistry, the pioneers (Coulson, Mac 
Weeny, just to quote two of them) used to solve fundamental equations modelling 
toy systems which could be simply numerically handled in view of their very limited 
size. The true difficulty arose with the need to model larger systems while possibly 
taking into account their interaction with their environment. Hand calculations 
were no longer possible, and computing science came into the picture. Today, the 
challenge is twofold: improving the formalism (both from a physical viewpoint and 
from the mathematical viewpoint), and speeding up rigorously founded numerical 
methods. From this originates the revival of interest in the communities of chemists 
and applied mathematicians. 
From what we heard from the audience of our series of minisymposia, the result 
of our enterprise was beyond our hope. So the idea carne out to translate that 
in a written manner. We therefore suggested that each of the nineteen lecturers 
should give a written account on his view on the subject. The topic of the written 
contribution need not be exactly the same as that of the talk, but. it. has to be 
related with the interplay between Mathematics, Numerical Analysis and Quantum 
Chemistry. The volume the two of us planned to edit was not a proceedings volume. 
It was rather an outgrowth of the series of talks, and an attempt to bring together 
this heterogeneous population this time in the written mode. The only constraint we 
imposed to the contributors was the following rule of the game: the mathematicians 
had to write in a language understandable by chemists, and vice-veT'.'ia. Thirteen 
lecturers out of ninet.een played the game until the end. The result is in the reader's 
1 ICIAM99, held in Edinburgh, Scotland, July 5-9th, 1999, 
x 
hands. It consists of eleven chapters devoted to various aspects of Computational 
Chemistry. 
This volume is divided into three parts. The first part deals with topics of general 
interest, the second one is devoted to the modelling of the condensed phases, the 
third one focuses on the relativistic aspects. 
The book opens with a contribution by Brian Sutcliffe on questions of symmetry 
in Quantum Mechanics. We find it symbolic that the first chapter of such a volume 
is written by such an eminent chemist, who is known to have had a constant interest 
into the mathematical aspects. 
The first part continues with a chapter by Eric Cances on the numerical analysis 
of SCF algorithms for HF calculations. Eric Cances is one of the representatives 
of this new generation of applied mathematicians who are very much involved in 
Computational Chemistry. He presents a rigorous analysis of the existing SCF 
algorithms and introduces new ones. 
Chapter 3 is written together by a chemist and a mathematician, Michel Caffarel 
and Roland Assaraf. It is devoted to Quantum Monte-Carlo methods. Their con-
tribution, that we consider in some sense as an instance of what should be done in 
order to enhance the links between the two communities, has been deliberately writ-
ten on a pedagogical tone. They are to be thanked for that (unfortunately unusual 
but so much useful) intention. 
The last chapter of the first part is due to Gabriel Turinici, who introduces the 
reader to the very important issue of exact control of quantum systems. 
As announced above, the second part of this volume deals with the modelling of 
the condensed phases. In Chapters 5 to 8, the crystalline solid phase is of concern. 
Chapter 9 is devoted to the liquid phase. 
Chapters 5 and 6 rather stand on the theoretical side. The former, due to 
Isabelle Catto, Pierre-Louis Lions and one of us (CLB), reports on a series of works 
devoted to the rigorous derivation of the models of the solid phase, a topic not so 
often addressed in the physical literature. The latter, written by Olivier Bokanowski, 
Benoit Grebert and Norbert Mauser, follows the same vein. It discusses the rigorous 
foundations of some well accepted approximations of Quantum Chemistry, such as 
the Xet method. 
In Chapter 7, Xavier Blanc, a young mathematician, presents his personal view 
on the models in use for the simulation of the crystalline phase. It may be interesting 
for a chemist expert at this topic to see the presentation of it by someone of the 
"outer world" . 
Chapter 8, from Vanina Louis-Achille and one of us (MD), is the logical sequel of 
Chapter 7; it develops the practical numerical aspects encountered in solid quantum 
chemistry calculations which are hidden in usual scientific papers. 
Chapter 9 by Benedetta Mennucci is an overview of liquid state methods of 
calculations. It provides the theoretical framework for the various methods currently 
used in Quantum Chemistry codes. 
The third part, devoted to the relativistic models, features two contributions. 
Xl 
Chapter 10 is due to Jean Dolbeault, Maria Esteban and Eric Sere. It reports on 
the works of the authors that give a sound mathematical ground to the Dirac-Fock 
model. By their contribution to the field, they bring the mathematical understand-
ing of the Dirac-Fock model to the level to that of the Hartree-Fock model. 
Trond Saue and H. J. Aa. Jensen introduce in Chapter 11 the various concepts 
of quaternion symmetry appearing in the Dirac Equations as they are implemented 
in the 4-component relativistic molecular calculations (DIRAC code). 
Let us end this foreword by emphasizing we strongly believe that the two com-
munities of applied mathematicians and chemists can take benefit of an interaction. 
At least we hope that our endeavour towards the challenge to bring together applied 
mathematicians and chemists will stimulate other ones. As we are tenacious, we are 
currently working on another project in the same vein. 
Needless to say, we encourage every interested reader to contact us. 
M. D. & C. L. B. 
Paris, April 2000. 
Part I 
General topics 
Chapter 1 
Is a molecule in chemistry 
explicable as a broken symmetry 
in quantum mechanics? 
Brian Sutcliffe 
Theoretical Division 1, 
Institute for Molecular Science, 
Myodaiji, Okazaki 444-8585, Japan, 
and present address 
Department de Chimie Physique Moleculaire, 
Universite Libre de Bruxelles, 
B-1050 Bruxelles, Belgium. 
bsutclif@ulb.ac.be 
Abstract : It is argued that incorporating the molecular geometry into the 
transformation specifications for deriving the standard (Eckart) Hamiltonian used 
to describe molecular spectra, cannot generallybe accomplished without breaking 
the permutational symmetry requirements on identical nuclei. 
M. Defranceschi et al., Mathematical Models and Methods for Ab Initio Quantum Chemistry
© Springer-Verlag Berlin Heidelberg 2000
4 Brian Sutcliffe 
1 Introduction 
The idea that the proper way to treat molecules in quantum mechanics is to treat 
their nuclear motions as essentially classical with the electrons treated quantum 
mechanically, dates from the very earliest days of the subject. The genesis of the 
idea is usually attributed to Born and Oppenheimer [1], but it is an idea that was 
in the air at the time, for the earliest papers in which the idea is used, predate 
the publication of their paper. The physical picture that informs the attempted 
separation is one well known and widely used even in classical mechanics, namely 
division of the problem into a set of rapidly moving particles, here electrons, and a 
much more slowly moving set, here the nuclei. Experience is that it is wise to try 
and separate such incommensurate motions both to calculate efficiently and to get 
a useful physical picture. 
The object of the separation in the molecular case is to get an electronic motion 
problem in which the nuclear positions can be treated as parameters and whose 
solutions can be used to solve the nuclear motion problem. The insights arising 
from classical chemistry seem to predicate that, for lowish energies, the nuclear 
motion function should be strongly peaked at a nuclear geometry that corresponds 
to the traditional molecular geometry. A function of this kind would allow a good 
account of the electronic structure of a molecule to be given in terms of a single 
choice for the nuclear geometry about which a molecule could be considered as 
performing small vibrations and undergoing essentially rigid rotations. Carl Eckart 
[2] was among the first to discuss how this picture might be supported. He did so 
in a context that assumed the separation of electronic and nuclear motions. In his 
approach, the electrons are regarded simply as providing a potential. This potential 
is invariant under all uniform translations and rigid rotations of the nuclei that form 
the molecule. It is usually referred to as a potential energy surface and the nuclei 
are said to move on this surface. Eckart actually treats the nuclear motions by 
classical rather than quantum mechanics but it is his approach and developments 
from it that have dominated the interpretation of molecular spectra since 1936. 
Because the approach is essentially classical, it is perfectly proper to treat identical 
particles as distinguishable, by virtue of their positions on the potential surface. 
This identification allows the assignment of a molecule to a point group, that of 
its equilibrium geometry, and such assignments have proved extremely fruitful in 
interpreting molecular vibration-rotation spectra. 
In what follows we shall not concern ourselves with the details of the separation 
of electronic from nuclear motion but rather with the consequences of trying to treat 
the nuclear motion quantum mechanically rather than classically and to allow for 
the permutational symmetry of identical nuclei and for the overall translation and 
rotation symmetries. 
Is a molecule in chemistry explicable as a broken symmetry... 5 
2 The equations of motion for the molecule 
Schrodinger's Hamiltonian describing the molecule as a system of N charged parti-
cles in a coordinate frame fixed in the laboratory is 
A !i2 N 1 2 e2 N I Z,Zj 
H(x)=--L-V (Xi)+- L-
2-1 mi 8HEa ·-1 Xi]· t- Z,)-
(1) 
where the separation between particles is defined by 
X7j = L(Xoj - Xoi)2 (2) 
o 
It is convenient to regard Xi as a column matrix of three cartesian components Xoi , 
a = X, y, z and to regard Xi collectively as the 3 by N matrix x. Each of the particles 
has mass mi and charge Zie. The charge-numbers Zi are positive for a nucleus and 
minus one for an electron. In a neutral system the charge-numbers sum to zero. 
To distinguish between electrons and nuclei, the variables are split up into two 
sets, one set consisting of L variables, xi, describing the electrons and the other set 
of H variables, xi, describing the nuclei and N = L + H. When it is necessary to 
emphasise this split, (1) will be denoted if (xn, xe). 
If the full problem has eigenstates that are square-integrable so that 
(3) 
then an eigenstate can be written approximately as a product of the required form 
The function, <I>(xn) (chosen to be non-trivial) is determined as a solution of an effec-
tive nuclear motion equation, and traditionally, a good guess for the electronic wave 
function is supposed to be provided by a solution of the damped nudei electronic 
Hamiltonian 
This Hamiltonian is obtained from the original one (1) by assigning the values ai 
to the nuclear variables x;', hence the designation clamped nuclei for this form. 
Within the electronic problem each nuclear position ai is treated as a parameter. 
For solution of the entire problem, the electronic wave function must be available for 
all values of these parameters. The energy obtained from the solution of this problem 
depends on the nuclear parameters and is commonly called the electronic energy. 
It is usual to think of the potential energy surface, used in the Eckart approach, as 
formed by adding the electronic and the classical nuclear repulsion energy. 
But it is not actually possible to use this approach directly to approximate so-
lutions, because the Hamiltonian (1) is invariant under uniform translations in the 
6 Brian Sutcliffe 
frame fixed in the laboratory. This means that the centre of molecular mass moves 
through space like a free particle and the states of a free particle are not quan-
tised and eigenfunctions are not square integrable. Before we consider this further 
let us note another two symmetries of the equation. The molecular Hamiltonian 
is invariant under all orthogonal transformations (rotation-reflections) of the par-
ticle variables in the frame fixed in the laboratory and is also invariant under the 
permutation of the variable sets of all identical particles. 
The way in which translational motion can be removed from the problem is 
well understood from classical mechanics. But it involves an essentially arbitrary 
choice of translation ally invariant coordinates. If we want to maintain a distinction 
between electronic and nuclear variables, it seems sensible to define L translation-
ally invariant electronic coordinates such that they transform into one another in the 
standard way under electronic permutations and are unaffected by nuclear permuta-
tions. The rotational motion should then be described in a translationally invariant 
way, entirely in terms of the original nuclear coordinates. This process consists es-
sentially of the choice of an orthogonal matrix C fixing a frame in the body of the 
molecule along with the choice of 3H - 6 internal coordinates qk which describe the 
changes in nuclear geometry and are invariant under any orthogonal transformation 
of the original coordinate set. The matrix C can be parameterised by three Euler 
angles which may be used to define the angular momentum of the system while 
the sign of detlCl specifies the handedness of the chosen frame. The coordinate 
transformation from the frame fixed in the laboratory to one fixed in the molecule 
is clearly a non-linear one and so must fail for some coordinate specification where 
the jacobian vanishes. So, for example, there will always be a molecular geometry 
at which, for any particular choice of embedding rule, it is not possible to define 
three Euler angles and here the transformation will fail. But there will always be 
a domain of validity and we shall confine attention to that domain. Within that 
domain the wave function forthe problem has the form 
+J 
T(XT ) L <I>~(q,z)IJMk > (5) 
k=-J 
where X T denotes the centre of mass coordinate. and q and z the internal nuclear 
and electronic coordinates respectively while IJ Mk > is an angular momentum 
eigenfunction depending on the three Euler angles. The translational motion can be 
completely separated off but the rotational motion can be separated only approx-
imately. If the separation between electronic and nuclear motion is valid then the 
internal motion function can be written in good approximation as 
(6) 
for any electronic state p and it is perfectly reasonable now to regard the electronic 
wave function as a solution of the clamped nuclei problem, a particular choice of the 
ai yielding a particular choice of the qk. 
A permutation of the variables describing a set of identical nuclei will naturally 
leave the centre of mass coordinate unchanged and therefore will not affect the re-
moval of translational motion. It will also leave the centre of nuclear mass coordinate 
Is a molecule in chemistry explicable as a broken symmetry... 7 
unchanged. But it is perfectly possible that such a permutation will induce changes 
in the orientation variables and in the internal coordinates and it might well be the 
case that it actually mixes the orientation variables and the internal coordinates. 
3 The Eckart choice of a frame fixed in the body 
In the Eckart approach to a fixing a frame in a molecule, the electrons are considered 
simply as providing a potential for nuclear motion and so in discussing this embed-
ding we shall ignore them. Implicitly however, the electronic coordinates are treated 
relative to the centre of nuclear mass so that the translationally invariant electronic 
coordinates are just like the original electronic coordinates in that there are just 
L of them, they are invariant under any permutations of the nuclei and transform 
in the standard way under electronic permutations. It is perhaps sufficiently clear 
from this that they cause no deep complications. 
The frame C fixed in the body is defined entirely in terms of the original nuclear 
coordinates and is such as to define a redundant set of cartesian coordinates z? 
according to 
x7 - X = Cz7 
where X is the centre of nuclear mass coordinate. Thus 
H 
Lmiz7 = 0 (7) 
i=1 
and the matrix C is chosen as to define a set of cartesians in the frame fixed in the 
body such that the reference structure is specified by 
where the ai are constant matrices and, by definition, 
H 
Lmiai = 0 
i=1 
The reference structure is, of course, usually chosen to be the classical molecular 
geometry. It is assumed chosen to reflect the equilibrium geometry of the molecule 
as it would be at the minimum of the potential. However for the purposes of carrying 
through the derivations, there is no need to specify it more closely than as a reference 
configuration. 
Defining the displacement coordinates as 
the specification of the matrix C is completed by requiring that 
H 
Lmiai x Pi = (5 (8) 
i=1 
8 Brian Sutcliffe 
provided that the ai do not define a line. If they do, then they provide only one 
rather than the required three constraints and the Euler angles of the Eckart frame 
cannot then be defined. Even when they do not define a line, it is possible that the 
displacement vectors might, and if they do, the definition of the Eckart frame again 
will fail. But we assume that this happens only for very large displacements from 
the reference molecular geometry and thus in a region where the vibrational wave 
function vanishes. 
These constraint conditions are on the components of the mass weighted sum 
over all the vectors, of the vector products of the reference geometry vectors with 
the displacement vectors. In classical mechanics the vanishing of these components 
would be interpreted as the system having no internal angular momentum at the 
reference geometry. The two conditions (7) and (8) defining the embedded frame are 
often called the Eckart conditions. The standard account of the Eckart formulation 
in quantum mechanical form is provided by Watson [3] and, for comparison, here 
zi is used for Watson's ri and ai for his r~. 
The zi are completely expressible in terms of a set of 3H - 6 internal coordi-
nates qk. The internal coordinates are expressed in terms of displacements from the 
reference geometry by 
H H 
qk = L L bexikPexi == L b~Pi' k = 1,2, ... 3H - 6 (9) 
i=l 0: i=1 
where the elements bexik are simply constants which may be regarded as components 
of a column matrix bik . The range of these coordinates is (-00, (0). 
Internal coordinates must be linearly independent and so the bik must be linearly 
independent and in order for an inverse transformation to exist between the internal 
coordinates and the coordinates in the frame fixed in the laboratory, it is [4] required 
that 
i=1 
Using the two Eckart conditions given above it follows that 
(10) 
where 
H 
B = L mi(xi' - X)a; (11) 
i=1 
The 3 by 3 matrix BTB is symmetric and therefore diagonalisable, so functions of it 
may be properly defined in terms of its eigenvalues. There are in principle, eight (23) 
possible distinct square root matrices, all such that their square yields the original 
matrix product. Consistency (see [5]) requires, however, that the positive square 
roots of each of the eigenvalues be chosen. The eigenvalues cannot be negative but 
one or more might be zero and if this is so then the Eckart frame cannot be properly 
defined This is a matter to which we shall return later. 
Is a molecule in chemistry explicable as a broken symmetry... 9 
If we denote the 3 by H matrix of the xi - X as wand the similar collection of 
all the zi and thus the ai as zn and a then, from (11), B may be written as 
where m is an H by H diagonal matrix with the nuclear masses along the diagonal. 
The second Eckart condition (8) can be manipulated to show that it is equivalent 
to the requirement that the matrix 
H 
A = L miziaf == znmaT (12) 
i=1 
is symmetric. Thus 
and so from (10) 
Whatever the precise specification made of the inverse square root, if A is singular 
then C is undefined and the Eckart specification fails. Thus, as mentioned above, 
if the ai together specify a line, for example if all the ayi and an vanish then A is 
clearly singular. But there is a special case. If the ai together specify a planar figure, 
for example if all the azi vanish, then A is again singular but in this case the Eckart 
conditions may be satisfied by requiring that the Z~i vanish too. The last row and 
column of A are null and there is a 2 by 2 non-vanishing block and, provided that 
this block is non-singular, the problem remains well defined. As a matter of fact 
there is no need in practice to treat this special case explicitly. The second Eckart 
condition implicitly orients the third axis to be perpendicular to the defined plane 
and, as long as the two by two block of A is non-singular, the planarity constraints 
are subsumed. 
It is possible to write the cartesians expressed in the Eckart frame directly in 
terms of the coordinates of the frame fixed in the laboratory as: 
and since the ij-th element of w T w is a scalar product, (xi - xyr (xj - X), the inter-
nal cartesians are clearly invariant under any orthogonal transformation, including 
inversion, of the original nuclear coordinates. 
3.1 The effects of nuclear permutations on the Eckart coor-
dinates 
If we represent a particular permutation of identical nuclei by P, where this is a 
standard orthogonal permutation matrix, then we can express any permutation of 
nuclei by 
xn ---t xnp 
10 Brian Sutcliffe 
and we see that this permutation induces in B the change 
B ---+ wpT maT = wm(aPf = Cznm(ap)T = CAP 
because the permutation is ofidentical nuclei, and so its representative matrix must 
commute with the mass matrix. Thus the permutation seems to act as if it changed 
the reference vectors but these are, of course, quantities fixed at the time of problem 
specification and hence unchanging parameters once a choice is made. From (10) it 
follows that under this permutation 
C ---+ CAP (APT Apr~ == CU 
If the matrix A P is singular then U will not exist, the change in C will be undefined, 
and hence the Eckart frame embedding will fail. So no such permutation can be 
realised in the Eckart formulation and a fundamental symmetry of of the problem 
is broken. 
If the permutation is such is that A P is non-singular and symmetric then it 
can be diagonalised by an orthogonal matrix V with eigenvector columns Vi and 
eigenvalues a;. In dealing with the square root, consistency again requires the 
choice of the positive value and U can therefore be written as 
3 
U = L sgn(af)v;v; 
;=1 
If all the a; are positive then U is just the unit matrix and if all negative then U 
is the negative unit matrix and represents an inversion. The form that it actually 
takes depends on the precise nature of the permutation. 
If U is a constant matrix, separation between angular and internal coordinates 
is essentially preserved. The most that can happen is that the internal coordinates 
go into linear combinations of themselves and the angular momentum components 
go into linear combinations of themselves. A special case of such a constant matrix 
is if the permutation is such that 
(13) 
then 
(14) 
The matrix S could arise as the orthogonal representation matrix of a point 
group operation on the molecular framework, for the point group of the molecu-
lar framework is isomorphic with a subgroup of the full permutation group of the 
molecule. The subgroup might be the full group, but that is only rarely the case. 
It is also the case a relation like (13) might be possible for a permutation or set 
of permutations that are unconnected with any point group operations. When the 
permutations can be realised by proper or improper rotation matrices, they are often 
called perrotations, a name introduced by Gilles and Philippot [6J. These matters 
and many other related ones, are discussed in more detail in Chapters 2 and 3 of 
Ezra's monograph [5]. 
In general the permutation will yield an unsymmetric A P but provided it is 
non-singular then U will be well defined and orthogonal. 
Is a molecule in chemistry explicable as a broken symmetry... 11 
An example 
As an example we shall choose ethene, which also illustrates the planar special case 
too: 
HI H4 
\ / 
CI C2 
/ \ 
H2 H3 
with the following choice of ai 
a, ~ ( -n a, ~ ( =n ~ ~ ( -J ) a, ~ ( n 
~~ cn~~ 0) 
Where al ... a4 represent the reference positions of the protons, each with mass mp 
and a5 and a6 the reference positions of the carbons, each with mass me. The system 
is specified to lie in the x - y plane and we shall be concerned only with the 2 by 
2 non-vanishing sub matrix of A. The point group of the reference figure is D 2h , 
using the Schonfliess notation which is standard in molecular spectroscopy. The 
full permutation group is S4 x S2 assuming the nuclei to be identical isotopes of 
hydrogen and of carbon respectively. The point group contains 8 operations while 
the permutation group contains 48 operations. 
It is convenient to rewrite A from (12) as 
where the split has been made into the proton and the carbon parts of the problem 
and permutations will occur only within each part. The carbon part of the problem 
is always of the form 
( ±mC(z~6 - z~5)d 0) = ±A o 0 c 
where the plus sign represents the identity permutation and the mmus SIgn the 
transposition of the carbons. If Ap is non-singular then 
and even though Ae is singular there is no reason to expect the second term on the 
right to be singular, except perhaps for particular values of the coordinates and we 
can assume that such regions of coordinate space may be treated as unvisited for 
our present purposes. It is possible that even if Ap is singular (Ap ± Ac) might not 
be, but this again will happen only in rather special regions. What has be(~n said 
12 Brian Sutcliffe 
in relation to Ap here, will hold equally for any A:. We would anticipate therefore 
that that any permutation of the protons that yields a non-singular A~' would yield 
a non-singular A P when taken with either of the possibilities for Ac. If the permuted 
form is symmetric, then A P will be symmetric, though there is no reason to expect 
this to be generally the case. With these matters in mind we shall concentrate on 
the proton permutations in what follows. 
Looking at the effect of the point group operations on the framework vectors it 
is easy to see that 
C2x == (jzx == (12)(34)(5)(6), C2y == (jyz == (14)(23)(56), C2z == i == (13)(24)(56) 
while (jxy is equivalent to the identity operation in the point group and the identity 
permutation. Thus the effect of these permutations on Ap is 
A~A(1 0) 
p p 0 -1 
c ( -1 0 ) Ap :2f Ap 0-1 
Adjoining the relevant transformed Ac to each of these matrices allows the constant 
matrix to be separated as S and we obtain precisely the results for U anticipated 
from (13) above. 
The proton permutations above are the complete set of double transpositions 
for the protons, and we would expect each of them, even when joined with the 
alternative carbon permutations to those of the point group, to yield non-singular 
A P . Hence the definition of the Eckart frame would be maintained even for such 
non-point group permutations. However one would not in general be able to factor 
out a common constant matrix for such permutations as one can for perrotations 
and one would expect the relevant U to be functions of the internal coordinates. 
It is difficult to analyse the problem further at this level of generality. However 
if we choose the reference values for the zi we can go a little further and if, with 
this special choice, the Eckart frame definition fails, then the frame certainly cannot 
be defined for small displacements, which form the most important region for the 
Eckart approach. For present purposes, therefore, we shall regard A as if composed 
amaT and the transformation as 
With these restrictions 
Ap = (4mt2 4~pb2) and Ac = (2mod2 ~) 
Looking at the six proton transpositions, four yield singular A: but two, (13) 
and (24), do not. The permutation (12) gives 
Is a molecule in chemistry explicable as a broken symmetry... 13 
and the matrix for (34) is the same. In neither case can the singularity be removed 
with the aid of Ac. The permutation (14) gives 
and the matrix for (23) is the same. In both cases the singularity can be removed 
with the aid of Ac. The permutation (24) yields 
and (13) is the same but with positive off-diagonal elements. Both matrices are 
non-singular and the Eckart frame can certainly be well defined in the case of either 
of these permutations associated with either of the carbon permutations. 
The eight permutations involving just three protons all lead to singular A; of 
the form 
so for none of these permutations can a satisfactory Eckart embedding be defined 
even when Ac is considered. 
Of the six permutations involving all four protons two, (1324) and (1423), yield 
singular A; of the form 
and two, (1243) and (1342), yield singular A; like 
This matrix can be rendered non-singular with the aid of Ac. The permutation 
(1234) yields the non-singular but unsymmetric A; 
and (4321) yields the negative of this matrix. 
Thus in the region of the reference geometry, of the 24 possible proton permu-
tations, 16 lead to singular A; and of these only 4 can lead to non-singular A P 
The remaining 8 proton permutations form a group and all lead tonon-singular A P 
combined with either of the possible carbon permutations. This group has itself a 
sub-group of order 4 which gives rise to the perrotations discussed above. 
14 Brian Sutcliffe 
3.2 Discussion and conclusions 
What emerges from what has been said above is precisely how the imposition of a 
classical molecular geometry as a reference structure on a system in order to sepa-
rate its internal motions from its rotations, breaks the permutational symmetry of 
the problem. It does so in such a way as to preclude a proper separation of rota-
tions and internal motions if full permutational symmetrisation is attempted. That 
a molecular structure seems plausible within a quantum mechanical context, is un-
doubtedly due to the role played by the solutions of the clamped nuclei Hamiltonian 
in considering the separation of electronic and nuclear motion. Indeed just this point 
was made many years ago in the first section of a very interesting paper by Berry 
[7]. But it is perhaps worthwhile pointing out that, providing that the electronic 
coordinates are chosen relative to the centre of nuclear mass, there is absolutely no 
reason why the Eckart conditions should not be imposed on the nuclear part of the 
full problem, without any explicit separation of electronic motion. It is our classical 
preconceptions about a molecule that cause the trouble, rather than anything to 
do with the perceived virtues or limitations of that separation usually attributed 
to Born and Oppenheimer. This observation, if correct, must limit too the validity 
(though not, of course the utility) of the notion of feasible permutations, introduced 
by Longuet-Higgins [8] to deal with this problem and so much developed by later 
workers (see, for example [9]). 
Since the problem is entirely of our own making it would seem foolish to persist 
in our error, particularly when there is an obvious way in which it can be avoided. If 
we choose the matrix C defining the rotational variables to be such as to diagonalise 
the instantaneous inertia tensor then all particles with the same mass will enter into 
the definition equivalently and permutational equivalence of the rotational variables 
will follow automatically. And indeed two of the first attempts to separate rotational 
from vibrational motion two, [10] and [11], adopted precisely this approach. It turns 
out that in this approach the definition of the frame embedded in the system depends 
upon the reciprocals of differences between instantaneous moments of inertia and 
therefore fails whenever two moments are equal. Thus the Hamiltonian could not 
be used to describe any symmetrical top molecule, for example ammonia. Indeed it 
turns out more generally, inspite of some heroic efforts by van Vleck [12], that the 
Hamiltonian so derived is largely ineffective in describing molecules in terms of their 
traditional geometrical structures and so it has found no use in the elucidation of 
molecular spectra. 
It seems therefore that at present molecular structure has to be incorporated 
into the Hamiltonian in order to describe spectra effectively but that the price for 
this incorporation is that of a fundamental symmetry breaking. 
Bibliography 
[1] M. Born and J.R. Oppenheimer, Ann.der Phys., 192784,457. 
[2] C. Eckart, Phys. Rev., 1935,47,552. 
[3] J.K.G. Watson, Mol. Phys., 1968, 15,479. 
[4] R. J. Malhiot and S. M. Ferigle, J.Chem. Phys, 22, 717-719, (1954). 
[5] G. Ezra, Symmetry properties of molecules, Lecture Notes in Chemistry 28, 
Springer-Verlag, Berlin, 1982. 
[6] J. M. F. Gilles and J. Philippot, Int. journ. quant. chem., 1972, 6, 225. 
[7] R. S. Berry, Rev. Mod. Phys., 1960, 32, 447. 
[8] H. C. Longuet-Higgins, Molec. Phys., 1963, 6, 445. 
[9] P. R. Bunker, Molecular Symmetry and Spectroscopy, Academic Press, Lon-
don, 1979. 
[10] C. Eckart, Phys. Rev., 1934., 46, 384. 
[11] J. O. Hirschfelder and E. Wigner, Proc. Nat. Acad. Sci., 1935,21, 113. 
[12] J. H. van Vleck, Phy". Rev., 1935,47,487. 
Chapter 2 
SCF algorithms for HF electronic 
calculations 
Eric Cand~s 
CERMICS, Ecole Nationale des Pants et Chaussees, 
6 fj 8, avenue Blaise Pascal, Cite Descartes, 
F-77455 Marne-La- Vallee Cedex, France 
cances@cermics.enpc.fr 
Abstract: This paper presents some mathematical results on SCF algorithms 
for solving the Hartree-Fock problem. In the first part of the article the focus is on 
two classical SCF procedures, namely the Roothaan algorithm and the level-shifting 
algorithm. It is demonstrated that the Roothaan algorithm either converges towards 
a solution to the Hartree-Fock equations or oscillates between two states which are 
not solution to the Hartree-Fock equations, any other behavior (oscillations between 
more than two states, "chaotic" behavior, ... ) being excluded. The level-shifting 
algorithm is then proved to converge for large enough shift parameter, whatever the 
initial guess. The second part of the article details the convergence properties of a 
new algorithm recently introduced by Le Bris and the author, the so-called Optimal 
Damping Algorithm (ODA). Basic numerical simulations pointing out the principal 
features of the various algorithms under study are also provided. 
M. Defranceschi et al., Mathematical Models and Methods for Ab Initio Quantum Chemistry
© Springer-Verlag Berlin Heidelberg 2000
18 Eric Cances 
1 Introduction 
The Hartree-Fock (HF) model is a standard tool for computing an approximation of 
the ground state of a molecular system within the Born-Oppenheimer setting. From 
a mathematical viewpoint, the HF model gives rise to a nonquadratic constrained 
minimization problem for the numerical solution of which iterative procedures are 
needed; such procedures are referred to as Self-Consistent Field (SCF) algorithms. 
The solution to the HF problem can be obtained either by directly minimizing the 
HF energy functional [6, 11, 17, 25] or by solving the associated Euler-Lagrange 
equations, the so-called Hartree-Fock equations [20, 21, 22]. 
SCF algorithms for solving the HF equations are in general much more efficient 
than direct energy minimization techniques. However, these algorithms do not a 
priori ensure the decrease of the energy and they may lead to convergence problems 
[23]. For instance, the famous Roothaan algorithm (see [21] and section 4) is known 
to sometimes lead to stable oscillations between two states, none of t.hem being a 
solution to the HF problem. This situation may occur even for simple chemical 
systems (see section 4). 
Many articles have been devoted to the important issue of the SCF convergence. 
The behavior of the Roothaan algorithm is notably investigated in [2, 12] and in 
[26, 27]. In [2, 12] convergence difficulties are demonstrated for elementary two-
dimensional models; in [26, 27], a stability condition of the Roothaan algorithm in 
the neighbourhood of a minimum of the HF energy is given for closed-shell sys-
tems. More sophisticated SCF algorithms for solving the HF equations have also 
been proposed to improve the convergence using various techniques like for instance 
damping [10, 28] or level-shifting [22]. Damping (as implemented in [28]) cures 
some convergence problems but many other remain. Numerical tests <confirm that 
the level-shifting algorithm converges towards a solution to the HF equations for 
large enough shift parameters; a perturbation argument is provided in [22] to prove 
this convergence in the neighborhood of a stationary point. Unfortunately there 
is no guarantee that the so-obtained critical point of the HF energy functional is 
actually a minimum (even local); in addition, the level-shifting algorithm is known 
to only offer a slow speed of convergence. In practice, the most commonly used SCF 
algorithm is at the present time the Direct Inversion in the Iteration Space (DIIS)algorithm [20]. Numerical tests show that this algorithm is very efficient in most 
cases, but that it sometimes fails. 
The present article belongs to a series of articles [3, 4] devoted to the SCF algorithms. 
Our first purpose here is to report on recent mathematical results on the conver-
gence properties of the Roothaan and of the level-shifting algorithms. Section 4 
concerns the Roothaan algorithm, which is the most "natural" algorithm for solving 
the HF equations. Its is demonstrated that the Roothaan algorithm either converges 
towards a solution to the HF equations or oscillates between two states which are 
not solution to the HF equations, any other behavior being excluded. This theo-
retical result is in accordance with the numerical experiments. It is then explained 
in Section 5 why the introduction of a "level-shift" makes the algorithm converge. 
SCF algorithms for HF electronic calculations 19 
The mathematical proofs are presented in the context of the finite dimension ap-
proximations of the HF problem obtained by a Galerkin method with a finite basis 
of atomic orbitals or plane waves, typically. They are consequently much simpler 
from a technical viewpoint than the proofs detailed in [3] which concern the original 
infinite dimension HF problem. 
Recently, new SCF algorithms have been introduced in [4] by Le Bris and the author. 
They seem to exhibit good convergence properties at least for the chemical systems 
computed so far. These algorit.hms have been called Relaxed Constraints Algorithms 
(RCA) for they can be interpreted as direct minimization procedure of the HF 
energy which do not care about satisfying at each it.eration the nonlinear constraints 
D2 = D that characterize admissible density matrices. The second purpose of this 
article (Section 6) is to detail the mathematical proof of the convergence of the basic 
RCA, namely the Opt.imal Damping Algorithm (ODA). Section 6 also contains some 
comments on the connections between RCA and other algorithms like the level-
shifting and the DIIS algorithms. 
Before coming up to our main topic, we devot.e Sect.ion 2 t.o a brief presentation 
of the HF model for readers (especially mathematicians) who are not familiar with 
Quantum Chemistry. Section 3 collects various general comments that apply to all 
the SCF algorithms considered in the sequel. 
2 A brief presentation of the Hartree-Fock model 
The problem under consideration consists in computing ab initio, that is to say 
without using any empirical parameter, the ground state energy of a molecular 
system made of M nuclei and N electrons. Tackling directly the M + N-body 
Schrodinger equation is today, and will probably remain, out of the scope of brute 
force numerical methods. Various approximations are therefore to be resorted to. 
The first approximation that is common to most models of Quantum Chemistry 
is the so-called Born-Oppenheimer approximation. To make short, it consists in 
considering the nuclei as classical point particles. The Born-Oppenheimer approx-
imation, which has been mathematically founded by Combes and al. [5], lays on 
the fact that nuclei are much heavier than electrons. The Born-Oppenheimer ap-
proximation is almost always valid in Chemistry (except for instance for studying 
specifically quantum phenomena involving nuclei as proton transfer by tunnel effect) 
and is therefore almost always used. 
Within the Born-Oppenheimer approximation, the searching for the ground state 
takes the form of two nested minimization problems: 
(1) 
with 
20 Eric Cances 
N 1 N M Zk 1 
H{xd = - L -~Xi - L L - + L 
i=l 2 i=l k=l IXi - Xk I l<'<J<N lx, - xJ I 
N 
1i = 1\ L 2 (R3 X {I+), I-)} , C) 
i=l 
In the above expressions, Xk denotes the current position in R3 of the k-th nucleus 
and Zk its charge. 
The N electrons are described by a wave function l/J( Xl, 0"1; ... ; X N, O"N), where 
Xi and O"i are respectively the position in 1R3 and the spin coordinate of the i-th 
electron. Each spin coordinate O"i can take two values here denoted by 1+) (spin up) 
and 1-) (spin down). The wave function l/J is a normalized vector of the fermionic 
Hilbert space 1i, and therefore satisfies on the one hand the antisymmetry condition 
l/J(Xp(l) , O"p(l);···; Xp(N), O"p(N)) = (-1)'(P)l/J(X1, 0"1;···; XN, O"N) 
for any permutation p of 1[1, N]I (c(p) denoting the signature of p), and on the other 
hand the normalization condition 
The operator H{xd is the so-called electronic hamiltonian. It acts on 1i; the Xk 
play the role of parameters. It is made of three terms, the first term accounting 
for the kinetic energy of the electrons, the second and the third terms accounting 
for nuclei-electrons and electrons-electrons interactions respectively. All physical 
quantities are expressed in atomic units [18]. 
Searching for the ground-state of the molecular system thus consists in minimizing 
the potential energy W (Xl, ... , X M) by solving the so-called geometry optimization 
problem (1). From the mathematical point of view, problem (1) is an unconstrained 
minimization problem of finite dimension. We refer the reader to [19, 24] for an 
overview of the various numerical methods dedicated to geometry optimization. 
The specificity of problem (1) is that the function to be minimized, namely the 
potential energy W, is itself the result (up to the inter-nuclear repulsion term 
L zkzL/lxk - xt!) of the minimization problem (2) which is usually referred to as 
the electronic problem. We face this time a constrained minimization problem on 
the infinite dimension space 1i. 
In the sequel, we focus on the electronic problem, which is rewritten (in order to 
simplify the notations) 
inf{ (l/J, Hl/J), l/J E 1i, Ill/JII = 1} (3) 
with 
N 
1i = /\ L 2 (R 3 X {I + ), 1-) }, C ) 
i=l 
SCF algorithms for HF electronic calculations 21 
N 1 N 1 
H = - L -6Xi + LV(Xi) + L 
;=1 2 i=1 1 <i<j<N IXi - Xj I 
M Zk 
V(x) = - L I - I 
k=1 X - Xk 
the Xk being now fixed parameters in lR3 . 
The Hartree-Fock approximation is of variational nature. It consists in restricting 
the set {1/! E 11., II1/! II = I} on which the energy functional (1/!, H1/!) is minimized 
to the set of the Slater determinants, i.e. to the set of the wave functions 1/! of the 
form 
(4) 
where the rPi, which are called molecular orbitals, satisfy the orthonormality condi-
tions 
A classical calculation (see [18] for instance) gives for any 1/) of the form (4) 
with 
N 
T<I>(X, 0"; x', 0"') = L rPi (x, 0") rPi(X', 0"')*, 
i=1 
N 
P<I>(x) = L L IrPi(X, O"W· 
i=l a 
The HF problem thus reads 
The mathematical properties of the HF problem have been studied by Lieb and 
Simon [14] and by P.-L. Lions [15]. The existence of a HF electronic ground state is 
guaranteed for positive ions (Z := L~I Zk > N) and neutral systems (Z = N). We 
are not aware of any general existence result for negative ions (the available existence 
proofs only work for N < Z + 1). On the other hand, there is a non-existence results 
for negative ions such that N > 2Z + M [13] (this inequality holds for instance for 
22 Eric CancEls 
the ion H2-). As far as we know, uniqueness (of the density p at least) is an open 
problem, probably of outstanding difficulty. 
The last step of the approximation procedure consists in approaching the infinite 
dimensional HF problem by a finite dimensional HF problem by means of a Galerkin 
approximation: the HF energy is minimized over the set of molecular orbitals that 
can be expanded on a given finite basis {xp} 1 < <n: 
-p-
n 
rPi = LCkiXk. 
k=1 
Denoting by S = [Sktl with 
the so-called overlap matrix, the constraints L k3 rPirP; = 6ij read 
cr 
or in matrix form 
C*SC=IN , 
where IN denotes the identity matrix of rank N. In addition, 
t G k31VrPd2 + k3 VlrPil2) 
~ (~ k31v ~ CikXkl2+ k3 V I~ CikX{) 
N n n 
L L L hklCliCki 
i=1 k=1 1=1 
Tr (hCC*) 
where h denotes the matrix of the core hamiltonian -~Ll + V in the basis {Xk}: 
Lastly, denoting by 
the inter-electronic repulsion term reads 
r r p.p(x) p.p(x') , r r IT.p(x, x')12 , n N * * 
if 3 if 3 I _ 'I dx dx - if 3 if 3 I _ 'I dx dx = L L AijklCioCje,ck/3Cl/3· 
R R X X R R X X i,j,k,I=10,/3=1 
SCF algorithms for HF electronic calculations 23 
The above expressions incite one to introduce the so-called density matrix 
D = CC*, 
which permits to write the HF energy under the compact form 
EHF(D) = Tr (hD) + ~Tr (G(D)D), 
where G(D) denotes the contracted product of the 4-index tensor A by D: 
G(D)ij = (A : D)ij = L AijklDkl. 
kl 
It is easy to see that the matrices D which read D = CC* with C E M(n, N) and 
C*SC = IN are those which satisfy Tr (SD) = Nand DSD = D. The so-obtained 
finite dimension HF problem then reads 
inf {EHF(D), DE M(n, n), D* = D, Tr (SD) = N, DSD = D}. 
For the sake of simplicity, we assume in the sequel that the overlap matrix S equals 
identity, that is to say that the basis {XpL<p<n is orthonormal. The general case is 
recovered by the transformation rules D -+- 51/ 2 DSI/2, h -+ S-1/2hS- l/2, G(D) -+ 
S-1/2G(Sl/2DS1/2)S-1/2. The HF problem then reads: 
(5) 
with 
p = {D E M(n, n), D* = D, Tr D = N, D2 = D}. 
The following lemma provides a characterization of the critical points of the HF 
minimization problem (5). 
Lemma 1. For any D, let us denote by F(D) = h+G(D) the Fock matrix associated 
with D. 
1. A density matrix DE P is a critical point of the HF problem (5) if and only 
if 
{ 
F(D)C = CE 
C'C = IN 
D=CC* 
(6) 
where E = Diag( 1'1,1'2, ... , EN) is a N x N diagonal matrix collecting N eigen-
values of the linear eigenvalue problem 
F(D)· ¢ = E¢. 
and where C is a n x N matrix containing N orthonor'mal eigenvectors asso-
ciated with 1'1, E2, ... , EN· The condition (6) is equivalent to the condition 
[F(D), D] = 0, 
where [".J denotes the matrix commutator defined for any A and Bin M(n, 71) 
by [A,B] = AB - BA. 
24 Eric Cand~s 
2. For D E P being a local minimum of the HF problem, it is necessary that (1, 
E2, ... , EN are the smallest N eigenvalues of F(D) including multiplicity. 
This result is classical; its proof can be read in any textbook of Quantum Chemistry 
(see [18] for instance). 
Remark. The model described above is the so-called General Hartree-Fock (GHF) 
model. Most often in practice, Quantum Chemistry calculations are performed with 
spin constraints models like the Restricted Hartree-Fock (RHF), the Unrestricted 
Hartree-Fock (UHF) or the Restricted Open-shell Hartree-Fock (ROHF) models. 
The convergence results stated below can be adapted without difficulties to these 
models. (; 
3 General remarks on SCF algorithms 
Various SCF algorithms are studied in the following three sections. All of them 
consist in generating a sequence (Dk ) defined by 
{ 
FkCk+1 = Ck+1 Ek+l 
Ck+1Ck+1 = IN 
Dk+l = Ck+l Ck+1 
(7) 
h E D· (k+l k+l) k+l < k+1 < < k+1 b . h· I were k+l = lag E1 , ... ,EN ,E1 _ E2 _ ... _ En emg t e e1genva ues 
of the linear eigenvalue problem 
and where Ck+1 collects N orthonormal eigenvectors associated with f~+l, E~+l, ... , 
Et+1. The expression of the current Fock matrix Fk characterizes the algorithm. We 
have for instance 
• Fk = F(Dk ) for the Roothaan algorithm; 
• Fk = F(Dk) -bDk where b is a positive constant for the level-shifting algorithm; 
• Fk = F(15k) for the ODA, where 15k is a pseudo-density matrix which satisfies 
the relaxed constraints 15~ ::::; Dk and is defined so that the HF energy EHF (15k) 
decreases at each iteration (see section 6). 
The procedure consisting in assembling the matrix D k+1 E P by populating the N 
molecular orbitals of lowest energies of the current Fock matrix Fk is referred to as 
the aufbau principle. It is justified by the results stated in Lemma 1. For the matrix 
Dk+1 being defined in a unique way, it suffices that Et+1 < Et+;l. Degeneracies in the 
spectrum are in general related to the symmetries of the system: in the cases when 
the system does not exhibit any symmetry, numerical experiments show that the 
eigenvalues of Fk are generically non-degenerate for any k, whereas it may not be the 
case when the system does exhibit symmetries (consider for instance the spherical 
SCF algorithms for HF electronic calculations 25 
symmetry of the hamiltonian in the atomic case). Degeneracies create technical 
difficulties which complicate the theoretical studies on SCF convergence. For the 
sake of simplicity, we therefore assume from now on that the uniform well-posedness 
(UWP) property introduced in [3] is satisfied: 
UWP property: a SCF algorithm of the form (7) with initial guess Do will be said 
to be uniformly well-posed if there exists some positive con8tant I such that 
The consequences of the UWP assumption which will be useful below have been 
collected in the following lemma, whose proof is postponed until the end of the 
present section. 
Lemma 2. Let us consider a SCF algorithm of the form (7) with initial guess Do 
which satisfies the UWP property. Then 
1. The updated density matrix Dk +1 is defined in a unique way at each iteration; 
this matrix can be characterized as the minimizer of the variational problem 
2. For any D E M(n,n) 8uch that D = D*, Tr (D) = Nand D2 S D, 
- - I 2 
Tr (FkD) ::: Tr (FkDk+d + 211D - Dk+111 , 
11·11 denoting the Hilbert-Schmidt norm defined for any A E M(n, n) by IIAII = 
Tr (AA*)1/2. 
In the sequel, we denote by arg inf MP the minimizer of the minimization problem 
MP. We can therefore write 
Dk+l = arg inf {Tr ChD), DE p}. 
Remark. Let us point out that some convergence results can be obtained without 
resorting to the UWP assumption. In particular, it turns out that the level-shifting 
algorithm is automatically UWP as soon as the shift parameter is large enough 
(see [3] for details). It can also be proved that the ODA numerically converges 
towards an aufbau solution to the HF equations within the GHF setting and provided 
the basis is "large enough". We do not detail here the rather technical proof of this 
assertion. Let us just mention that it is based on a mathematical result by P.-L. 
Lions [16] related to finite-temperature HF models. Unfortunately, so far as we 
know, the arguments used in [16] cannot be extended to the RHF, L"HF or ROHF 
models. <> 
Before turning to the study of SCF algorithms, the notion of convergence has to be 
made precise. We are in fact not able to prove mathematical convergence results of 
26 Eric Cances 
the form "the sequence (Dk) converges towards a minimizer D of the HF problem 
(5)" for at least two reasons. First, we are solving the Euler-Lagrange equations 
associated with the HF minimization problem (5), namely the HF equations (6); 
even in case of convergence we have no argument to conclude that the so-obtained 
critical point is actually a minimum (even local) of the HF energy. Second, we have 
no precise description of the topology of the set of the critical points of (5); this lack 
of information prevents us from proving the convergence of the whole sequence (Dk ) 
towards a solution D to the HF equations. We can at best obtain that Dk+l - Dk 
goes to zero, and that for "large" k, Dk is "close to" a solution to the HF equations 
(6) satisfying the aufbau principle. For instance, it may happen that the HF problem 
admits a connected manifold of minima; this phenomenon is observed in particular 
for open-shell atoms because the spherical symmetry of the problem is broken by the 
HF approximation (this can be related to a mathematical result by Bach, Lieb, Loss 
and Solovej [1] stating that "there are no unfilled shell" in the HF ground states). 
We cannot then discriminate between the casewhen the sequence (Dk ) converges 
towards a point of the manifold and the case when the sequence (Dk ) is attracted 
by the manifold together with a slow drift parallel to the manifold. 
We shall consequently adopt here the following two convergence criteria, which are 
sufficient in practice. We shall say that a SCF algorithm of the form (7) numerically 
converges towards a solution to the HF equations if the sequence (Dk) satisfies 
and that it numerically converges towards an aufbau solution to the HF equations 
if the sequence (Dk) satisfies 
As all norms are equivalent in finite dimension, we do not need to specify the ma-
trix norm in which the variations are evaluated. Let us remark that the latter 
convergence criterion is stronger than the former one for 
Let us conclude this section with the 
Proof of Lemma 2. Let us denote by D a current matrix such that D = D*, 
Tr (D) = N, D2 ::; D and by Dij its coefficients in an orthonormal basis in which 
Fk = Diag( f~+l, f~+l, ... , f~+l) with f~+l ::; f~+l ::; '" ::; f~+l. In such a basis 
Dk+1 = Diag(l, .. ·,1,0, ... ,0). As in addition Tr (D) = I:?=1 Dii = N, we get first 
Tr ((Dk+1 - D)· (Dk+1 - D)) 
Tr (m+l) + Tr (D2) - 2 Tr (DDk+1) 
SCF algorithms for HF electronic calculations 27 
< Tr (Dk+l ) + Tr (D) - 2Tr (DDk+d 
N 
2N - 2LDii 
i=1 
n 
= 2 L D ii · 
i=N+l 
0:::; Dii :::; 1, for any 1 :::; i :::; n 
for D2 :::; D = D* implies IDid2 + LJt'i IDiJI2 :::; Dii . Putting together the above 
results, we obtain 
n 
L E7+1Dii 
i=1 
N n 
> L E7+1Dii+ L (Et+I+,)Dii 
i=1 i=N+I 
N n n 
L E7+ 1 Dii + ft+ 1 L Dii +, L D'i 
i=1 i=N+l i=N+1 
N N n 
L E7+ 1 Dii + Et+I(N - L Dii ) +, L Dii 
i=1 i=1 i=N+l 
N N n 
L E7+l + L(Et+ 1 - E7+ 1)(1- D ii ) +, L D i ,. 
i=1 i=1 
As for any 1 :::; i :::; N, 0 :::; Dii :::; 1 and Et+1 2: E7+ 1, we finally obtain 
- - , 2 
Tr (FkD) 2: Tr (FkDk+d + 2" liD - Dk+1 11 . 
The two statements of Lemma 2 follow. <) 
4 The Roothaan algorithm: why and how it fails 
The Roothaan algorithm (also called simple SCF or pur'e SCF or conventional SCF 
in the literature) is the simplest fixed point procedure associated with the nonlinear 
eigenvalue problem (6). It consists in generating a sequence (D:th) in P satisfying 
{ 
F(D:th)Ck+l = Ck+1Ek+1 
Ck+I Ck+1 = IN 
DRth C C* k+l = k+l k+l 
h E D· (k+l k+l) k+l < k+l < < k+l b' h M II were k+l = lag El ,"',EN , EI _ E2 _'" _ EN elllg t e jV sma est 
eigenvalues of the linear eigenvalue problem 
28 Eric Cand~s 
and where the n x N matrix CHI collects N orthonormal eigenvectors of F(Dfth) 
associated with (~+I, (~+1, ... , (~+1. The iteration procedure of the Roothaan algo-
rithm can therefore be summarized by the diagram 
DRth k+l· 
The convergence properties of the Roothaan algorithm are not satisfactory: although 
the Roothaan algorithm sometimes numerically converges towards a solution to the 
HF equations, it frequently numerically oscillates between two states, none of them 
being solution to the HF equations. Numerical oscillation between two states means 
here that 
DRth _ DRth --+ 0 
k+2 k , but DRth _ DRth ~ 0 k+1 k -,r . 
The behavior of the Roothaan algorithm can be explained by introducing the aux-
iliary function 
E(D, D') = Tr (hD) + Tr (hD') + Tr (G(D) D'), 
which is symmetric since Tr (G(D) D') = Tr (G(D') D), and which satisfies E(D, D) = 
2 EHF(D). Let us indeed minimize E alternatively with respect to each of the two 
arguments D and D': 
DI=arginf{E(Do,D), DEP}, 
D2 = arg inf{E(D, DI), DE P}, 
D3 = arg inf {E(D2' D), DE P}, 
This minimization procedure is usually called relaxation in the mathematical liter-
ature. For the first two steps, we obtain 
Dl arg inf {E(Do, D), DE P} 
arg inf {Tr (hDo) + Tr (hD) + Tr (G(Do)D), DE P} 
arg inf {Tr (F(Do)D), D E P} 
DRth 
1 , 
and, since E is symmetric on P x P, 
D2 arg inf {E(D, Drth ), DE p} 
arg inf {E(Drth, D), D E p} 
arg in£{ Tr (hD) + Tr (hDfth) + Tr (G(Drth)D), DE p} 
arg in£{ Tr (F ( Dfth ) D) , D E p} 
Dfth. 
SCF algorithms for HF electronic calculations 29 
It follows by induction that the sequences generated by the relaxation algorithm on 
the one hand, and by the Roothaan algorithm on the other hand, are the same. 
The functional E, which decreases at each iteration of the relaxation procedure can 
therefore be interpreted as a Lyapunov functional of the Roothaan algorithm. This 
basic remark is the foundation of the proof of the following result. 
Theorem 1. Let Do E P such that the Roothaan algorithm with initial guess Do is 
UWP. Then the sequence (Dfth) generated by the Roothaan algorithm satisfies one 
of the following two properties 
• either (Dfth) numerically converges towards an aufbau solution to the HF equa-
tions 
• or (Dfth) numerically oscillates between two states, none of them being an 
aufbau solution to the HF equations. 
Proof. For any kEN, we deduce from Lemma 2 that 
Tr (F(DRth)DRth) + 211DRth - DRthl12 < Tr (F(DRth)DRth) k+1 k+2 2 k+2 k - k+l k . 
Adding Tr (hDt:!1) to both terms of the above inequality, we obtain 
E(DRth DRth) + 211DRth _ DRthl12 < E(DRth DRth) 
k+1' k+2 2 k+2 k - k' k+l' 
We then sum up the above inequalities for kEN and we get LkEN IIDf~1- Dfth 112 < 
+00, which involves in particular that 
DRth _ DRth --+ 0 k+2 k . 
Now, either Df~1 - Dfth converges to zero or it does not. In the former case, we 
deduce from the characterization of Df~1 by 
that 
Tr (F(Dfth)Dfth) - inf {Tr (F(Dfth)D), DE p} --+ O. 
Convergence towards an aufbau solution to the HF equations is thus established. In 
the latter case 
Tr (F(D:~h)D:~h) - inf {Tr (F(D:~h)D), DE p} 
= Tr (F(D:~h)D:~h) - Tr (F(D:~h)D:~~l) 
> 211DRth _ DRth 112_Jc O. 
- 2 2k 2k+l--r' 
Convergence of (D2k) towards an aufbau solution to the HF equations is therefore 
excluded; the same argument holds for (D2k+d· <> 
30 Eric Cances 
Mimicking the proof of Theorem 2 (see section 5), it is easy to establish in addition 
that (D2k> D2k+1) converges up to an extraction to a critical point (D, D') E P x P 
of the functional E which satisfies 
F(D')C = CE 
C'C = IN 
D=CC' 
F(D)C' = C'E' 
C"C' = IN 
D' = C'C'* 
where E and E' are diagonal matrices collecting the smallest N eigenvalues of F(D') 
and F(D) respectively. Besides, as E(D2k' D2k+1) is decreasing, the whole sequence 
(D2k' D2k+1) converges to (D, D') if this critical point is a strict (local) minimum. 
In this case, the alternatives are 
• either (D, D') is on the diagonal of P x P (i.e. D = D') and (Dfth) converges 
towards an au/bau solution to the HF equations; 
• or (D, D') is not on the diagonal of P x P (i.e. D #- D') and (Dfth) oscillates 
between two states which are not au/bau solutions to the HF equations. 
Both situations are represented on Figure 2.1. 
0' 0' , 
......-::--:::-:---.F----- ~ D __________ ~--~~__.~ D 
Figure 2.1: Minimization of E by relaxation: convergence towards a strict local 
minimum located on (resp. off) the "diagonal" leads to the convergence (resp. 
oscillations) of the Roothaan algorithm. 
Oscillations can be observed even for simple chemical systems. As a matter of ex-
ample, we have tested the Roothaan algorithm in the UHF setting for the atoms 
of the periodic table and for two sets of atomic orbitals, namely the Gaussian ba-
sis sets 3-21G and 6-311++G(3df,3pd) (see [9]). The initial guess is obtained by 
diagonalization of the core hamiltonian. Calculations have been performed with 
Gaussian 98 [8]. The results are reported in Figure 2.2; they indicate that 
SCF algorithms for HF electronic calculations 31 
1. Both alternatives (convergence vs oscillation) are met in practice. 
2. Convergence towards a critical point of the HF problem which is not a global 
minimum can sometimes be observed. 
3. For the same system, we can getconvergence for one basis set and oscillation 
for another basis set. 
111111111111111 111111111111111 
111111111111111 111111111111111 
......................................... 
• 
._ ........ __ 10_ .. ',-
----
o __ ..-IoC-,. 
Figure 2.2: Searching the ground state of atoms with the Roothaan algorithm. 
Results are shown on the periodic table of the elements. 
5 Level-shifting 
The analysis developed in the previous section suggests to add to the functional E 
a penalty term Ep of the off-diagonal pairs (D, D') with D # D' in order to enforce 
the critical points of the functional E + Ep to lie on the diagonal of P x P, which 
should ensure convergence towards a critical point of (5). 
A simple penalty fonctional reads Ep = biiD - D'11 2 , where b is a positive constant 
and where II . II denotes as above the Hilbert-Schmidt norm. Let us therefore set 
Eb(D, D') = Tr (hD) + Tr (hD') + Tr (G(D) D') + b liD - D'I1 2 
The relaxation algorithm associated with the minimization problem 
in£{ Eb(D, D') , (D , D') E P x p} 
32 Eric Cances 
generates the sequence (D~) defined by 
aufbau Db 
---+ k+1' 
The sequence (Dn can be identified with the sequence generated by the so-called 
level-shifting algorithm [22] with level-shift parameter b. The convergence of the 
level-shifting algorithm towards a (non necessarily aufbau) solution to the HF equa-
tions is mathematically guaranteed: 
Theorem 2. There exists a positive constant bo such that for any Do E P and for 
any level-shift parameter b 2: bOl 
1. The sequence of the energies EH F (D~) decreases towards some stationary value 
E of EHF. 
2. The sequence (D~) numerically converges towards a solution to the HF equa-
tions. 
Proof. Let bo be a positive constant such that 
V(D, D') E M(n, n) x M(n, n), Tr (G(D - D') . (D - D')) :::; bollD - D'11 2 . (8) 
Such a bo exists since (d, d') f-7 Tr (G (d)d') is a bilinear form on M (n, n). As Eb is 
symmetric on P x P, we have for any k E IN, 
Eb(DZ, DZ+1) in£{ Eb(DZ, D), DE p} 
< Eb(D%, DZ). 
A simple calculation shows that this inequality can be rewritten as 
Therefore, for any b 2: bo, 
(9) 
It follows that (EHF(D~)) is a decreasing sequence and that 
+00 
L IID%+l - DtW < +00. 
k=O 
The latter statement, which has been obtained by summing the inequalities (9) for 
k 2: 0, implies in particular that 
As for any k E IN, 
SCF algorithms for HF electronic calculations 33 
it follows that 
This concludes the proof of statement 2. As P is compact, we can extract from 
(D%)kEN a subsequence (Dt)IEN which converges towards some D E P, such that 
EHF(D%) -I- EHF(D) and [F(D),D] = 0; E. = limEHF(Dk) = EHF(D) is therefore 
a stationary value of the HF energy. <) 
The level-shift parameter bo implicitly defined by (8) is far from being optimal. Ex-
plicit and more refined estimates of shift parameters that guarantee convergence are 
given in [3]. From a numerical viewpoint, it is important to choose not too large a 
shift parameter; otherwise the speed of convergence is very slow and the risk of con-
verging towards a critical point whose energy is above that of the HF ground state is 
enhanced. This point, on which we will come back in the course of section 6, is illus-
trated by the numerical example reported on Figure 2.3 in which the level-shifting 
algorithm with various shift parameters has been used to compute the (doublet) 
UHF ground state of the Bromine atom in the Gaussian basis 6-311 ++G(3df,3pd) 
(see [9]). In each case, the initial guess is computed by diagonalizing the core hamil-
tonian. Calculations have been performed with Gaussian 98 [8]. The algorithm 
oscillates for small shift parameters. For larger shift parameters, damped oscilla-
tions leading to convergence are observed. For very large shift parameter, the energy 
decreases at each iteration. Too large shift parameters have however to be excluded 
because they slow down the convergence (for b = 30.0 Ha, convergence towards the 
ground state is obtained after more than 200 iterations). 
6 The Optimal Damping Algorithm 
The present section is devoted to the mathematical study of the Optimal Damp-
ing Algorithm (ODA) which is the simplest representative of the class of Relaxed 
Constraints Algorithms (RCA) introduced in [4]. 
The ODA is defined by the following two-step iteration procedure 
l. Diagonalize the current Fock matrix Fk = F(i5k) and assemble the matrix 
Dk +1 E P by the aufbau principle; 
denotes the line segment linking Dk and Dk +1 . 
The procedure is initialized with Do = Do, Do E P being a given initial guess. 
The ODA thus generates two sequences of matrices: 
34 Eric Cances 
.-
-
L .. r 
• 
, 
.-
.. .. .. .. .. .. .. .. 
-
1- :;::.:; ~:.~ 1 r-· ::=:;!!.--l 
.... 
-
1- ~ 1-
_\ - \ -
c 
.. .. .. .. .. 
-
I .. :=:;!:.. ] 
,-
I_ I 
• • 
.-
.. .. .. .. .. .. .. • .. 
-
Figure 2.3: Calculation of the UHF ground state of the Bromine atom . 
• The principal sequence of density matrices (Dk)kEN which will be proved to 
numerically converge towards an aufbau solution to the HF equations; 
• A secondary sequence (Dkh?l of pseudo-density matrices which belong to the 
set 
P = {15 E M(n,n), Tr (D) = N, 
obtained from P by relaxing the nonlinear constraints D2 = D. 
The latter statement is a direct consequence of 
Lemma 3. The set P is convex, 
whose proof is postponed until the end of the present section. Indeed, Do = Do E 
PcP and, by ind~lction, if 15k E P then by convexity Dk+1 E Seg[Dk, D k+1] C P 
since D k+1 E PCP. 
SCF algorithms for HF electronic calculations 35 
The properties of the ODA are put together in the following theorem. 
Theorem 3. For any initial guess Do for which the aDA is UWP, 
1. The sequence E(i5k ) decreases towards a stationar'y value of the HF energy. 
2. The sequence (DkhEN converges towards an aufbau solution to the HF equa-
tions. 
As a first step towards the understanding of the ODA, let us consider 15k E 15 and 
D' E P, and let us compute the variation of the HF energy on the line segment 
Seg[15k , D'] = {(I - A)15k + AD', A E [0, I]}. 
We obtain for any A E [0,1]' 
EHF((l - A)15k + AD') 
= EHF (15k) + ATr (F(15k) . (D' - 15k)) + ~2 Tr (G(D' - 15k) . (D' - 15k)) . 
The "steepest descent" direction, i.e. the density matrix D for which the slope 
Sjjk-+ D = Tr (F(15k) . (D - 15k)) is minimum, is given by the solution to the mini-
mization problem 
which also reads 
D = arg inf {Tr (F(Dk) . D'), D' E.P} . 
This is precisely the direction Dk+1 obtained by the aufbau principle. The ODA 
can therefore be interpreted as a steepest descent algorithm in 15. The practical 
implementation of the ODA is detailed in [4] for the RHF setting. The cost of 
one ODA iteration is approximatively the same as the cost of one iteration of the 
Roothaan algorithm (see [4] for details). 
Figure 2.4 reports on a comparison between the ODA and the DIIS approaches for 
the calculation of the RHF ground state of the E form of n-methyl-2-nitrovinylamine 
(CH3-NH-CH=CH-N02 ) in the basis 6-31G(d) (see [7]). The speed of convergence is 
estimated by computing the logarithm of the difference between the HF energy of the 
current density matrix and the (presumed) HF ground state energy. Calculations 
have been performed within Gaussian 98 [8]. The graph on the left hand side 
corresponds to an initial guess computed by a semi-empirical method. In this case, 
both algorithms converge but the speed of convergence of the DIIS algorithm is 
higher. From a general viewpoint, numerical tests performed until now demonstrate 
that the ODA is efficient for performing the early iterations of the SCF procedure; 
when the sequence (Dk ) has reached a neighbourhood of a critical point of the HF 
problem, convergence can be accelerated either by resorting

Continue navegando