2018 Book ConceptsMethodsAndApplications
2018 Book ConceptsMethodsAndApplications
Yan A. Wang
Mark Thachuk
Roman Krems
Jean Maruani Editors
Concepts, Methods
and Applications of
Quantum Systems in
Chemistry and Physics
Selected Proceedings of QSCP-XXI
(Vancouver, BC, Canada, July 2016)
Concepts, Methods and Applications of
Quantum Systems in Chemistry and Physics
Progress in Theoretical Chemistry and Physics
VOLUME 31
Honorary Editors
Rudolph A. Marcus (California Institute of Technology, Pasadena, CA, USA)
Roy McWeeny (Università di Pisa, Pisa, Italy)
Editors-in-Chief
J. Maruani (formerly Laboratoire de Chimie Physique, Paris, France)
S. Wilson (formerly Rutherford Appleton Laboratory, Oxfordshire, UK)
Editorial Board
E. Brändas (University of Uppsala, Uppsala, Sweden)
L. Cederbaum (Physikalisch-Chemisches Institut, Heidelberg, Germany)
G. Delgado-Barrio (Instituto de Matemáticas y Física Fundamental, Madrid, Spain)
E. K. U. Gross (Freie Universität, Berlin, Germany)
K. Hirao (University of Tokyo, Tokyo, Japan)
Chao-Ping Hsu (Institute of Chemistry, Academia Sinica, Taipei, Taiwan)
R. Lefebvre (Université Pierre-et-Marie-Curie, Paris, France)
R. Levine (Hebrew University of Jerusalem, Jerusalem, Israel)
K. Lindenberg (University of California at San Diego, San Diego, CA, USA)
A. Lund (University of Linköping, Linköping, Sweden)
M. A. C. Nascimento (Instituto de Química, Rio de Janeiro, Brazil)
P. Piecuch (Michigan State University, East Lansing, MI, USA)
M. Quack (ETH Zürich, Zürich, Switzerland)
S. D. Schwartz (Yeshiva University, Bronx, NY, USA)
O. Vasyutinskii (Russian Academy of Sciences, St Petersburg, Russia)
Y. A. Wang (University of British Columbia, Vancouver, BC, Canada)
Concepts, Methods
and Applications of
Quantum Systems in
Chemistry and Physics
Selected Proceedings of QSCP-XXI
(Vancouver, BC, Canada, July 2016)
123
Editors
Yan A. Wang Roman Krems
University of British Columbia University of British Columbia
Vancouver Vancouver
Canada Canada
This Springer imprint is published by the registered company Springer International Publishing AG
part of Springer Nature
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
PTCP Aim and Scope
v
vi PTCP Aim and Scope
such as mathematical chemistry (which involves the use of algebra and topology in
the analysis of molecular structures and reactions); molecular mechanics, molecular
dynamics and chemical thermodynamics, which play an important role in rational-
izing the geometric and electronic structures of molecular assemblies and polymers,
clusters and crystals; surface, interface, solvent and solid state effects; excited-state
dynamics, reactive collisions and chemical reactions.
Recent decades have seen the emergence of a novel approach to scientific
research, based on the exploitation of fast electronic digital computers. Computation
provides a method of investigation which transcends the traditional division
between theory and experiment. Computer-assisted simulation and design may
afford a solution to complex problems which would otherwise be intractable to
theoretical analysis, and may also provide a viable alternative to difficult or costly
laboratory experiments. Though stemming from theoretical chemistry, computa-
tional chemistry is a field of research in its own right, which can help to test
theoretical predictions and may also suggest improved theories.
The field of theoretical molecular sciences ranges from fundamental physical
questions relevant to the molecular concept, through the statics and dynamics of
isolated molecules, aggregates and materials, molecular properties and interactions,
to the role of molecules in the biological sciences. Therefore, it involves the
physical basis for geometric and electronic structure, states of aggregation, physical
and chemical transformations, thermodynamic and kinetic properties, as well as
unusual properties such as extreme flexibility or strong relativistic or quantum field
effects, extreme conditions such as intense radiation fields or interaction with the
continuum, and the specificity of biochemical reactions.
Theoretical chemistry has an applied branch (a part of molecular engineering),
which involves the investigation of structure–property relationships aiming at the
design, synthesis and application of molecules and materials endowed with specific
functions, now in demand in such areas as molecular electronics, drug design or
genetic engineering. Relevant properties include conductivity (normal, semi- and
super-), magnetism (ferro- and ferri-), optoelectronic effects (involving nonlinear
response), photochromism and photoreactivity, radiation and thermal resistance,
molecular recognition and information processing, biological and pharmaceutical
activities, as well as properties favoring self-assembling mechanisms and combi-
nation properties needed in multifunctional systems.
Progress in Theoretical Chemistry and Physics is made at different rates in these
various research fields. The aim of this book series is to provide timely and in-depth
coverage of selected topics and broad-ranging yet detailed analysis of contemporary
theories and their applications. The series will be of primary interest to those whose
research is directly concerned with the development and application of theoretical
approaches in the chemical sciences. It will provide up-to-date reports on theoretical
methods for the chemist, thermodynamician or spectroscopist, the atomic, molecular
or cluster physicist, and the biochemist or molecular biologist who wish to employ
techniques developed in theoretical, mathematical and computational chemistry in
their research programs. It is also intended to provide the graduate student with a
readily accessible documentation on various branches of theoretical chemistry,
physical chemistry, and chemical physics.
Preface
This volume collects 20 selected papers from the scientific contributions presented
at the Twenty-first International Workshop on Quantum Systems in Chemistry,
Physics, and Biology (QSCP-XXI), organized by Yan Alexander Wang at the
University of British Columbia in Vancouver, BC, Canada, on July 02–09, 2015.
Over 160 scientists from 30 countries attended this meeting. The participants dis-
cussed the state of the art, new trends, and future evolution of methods in molecular
quantum mechanics and their applications to a broad variety of problems in
chemistry, physics, and biology.
The high-level attendance attained in this conference was particularly gratifying.
It is the renowned interdisciplinary nature and friendly feeling of QSCP meetings
that make them so successful discussion forums.
Highly ranked among the world best universities, the University of British
Columbia (UBC) holds an international reputation for excellence in advanced
research and teaching. Only 30 min from the vibrant heart of downtown Vancouver,
the spectacular UBC campus is a ‘must-see’ for any visitor in the world: Snow-
capped mountains can be seen meeting the ocean, and breathtaking vistas greet you
around every corner. The UBC campus also boasts some of the city’s best attrac-
tions and recreation facilities, including the Chan Centre for the Performing Arts,
the Museum of Anthropology, the UBC Botanical Garden, and endless opportu-
nities to explore forested trails in the adjoining 763-ha Pacific Spirit Regional Park.
Details of the Vancouver meeting, including scientific and social programs, can be
found on the Web site: https://groups.chem.ubc.ca/qscp/. Altogether, there were 30
morning and afternoon sessions, where 48 plenary lectures and 36 invited talks were
given, and one evening poster session, with 18 posters being displayed. We are grateful
to all 166 participants for making the QSCP-XXI workshop a stimulating experience
and a great success. QSCP-XXI followed the traditions established at previous
workshops:
QSCP-I, organized by Roy McWeeny in 1996 at San Miniato (Pisa, Italy);
QSCP-II, by Stephen Wilson in 1997 at Oxford (England);
vii
viii Preface
The Editors
Contents
Preface vii
xi
xii Contents
Bing Wang, Paul A. Dobosh, Stuart Chalk, Keigo Ito, Mirek Sopek
and Neil S. Ostlund
Abstract Chemical Semantics, Inc. (CSI) is a new start-up devoted to bringing the
Semantic Web to chemistry and biochemistry. The semantic web is referred to as
Web 3.0 or alternatively the Web of Data or the Web of Meaning. It does not
replace the existing World Wide Web but augments it, placing data on the web in a
structured form such that the data has “meaning” and computers can understand it.
CSI has created a demonstration portal for exploring this new technology, specif-
ically at this point for data created by quantum chemistry calculations. This paper
describes the basics of a semantic web portal and the fundamental technology we
have used in developing it.
Data on the existing World Wide Web is encoded in documents containing text,
table, images, etc. This data is not really recognized by computers but has to be
interpreted by humans. Because computers cannot recognize the data in existing web
documents, the data cannot properly be shared or even found. The semantic web [1]
puts data onto the web in a form that allows computers to properly recognize it along
with its meaning. Computers can then perform intelligent operations on the data,
ultimately creating new data by using inference from existing data. The future of
scientific data involves structuring the data via the semantic web or equivalent
technologies so that computers can find data and understand data on our behalf.
Chemistry generates enormous amounts of data [2]. Because existing publication
channels do not give credence to the data in the way they give credence to the text
of a “scientific publication” much of this data is lost or discarded and not made
available to other scientists. There is a trend towards journals requiring authors to
submit data files along with the text of a publication. However, since there is as yet
no standard and little infrastructure adopted by mainstream publishers for dealing
with this data, most of it remains in inaccessible files (e.g. PDF documents). An
appropriate answer to this dilemma is the semantic web.
In conjunction with the data, the semantic web includes a vocabulary for
describing the data. This is where semantics comes to the fore. This vocabulary for
a field such as Quantum Chemistry needs to be encoded in a formal language. The
Web Ontology Language (OWL) [3] is such a language and the “vocabulary” is
really an ontology describing the formal language of a specific domain such as
computational or quantum chemistry. Chemical Semantics has created such an
ontology which is referred to as the Gainesville Core (http://purl.org/chem/gc). This
first ontology for quantum chemistry will need to be modified and extended by
scientists in the field as basic ontology ideas become more prevalent in chemistry.
Our ontology is meant to be placed in the public domain so that the quantum
chemistry community can both contribute to it and use it. Developing the Gai-
nesville Core is an ongoing project with release numbers.
The semantic web allows scientists to publish their data in a structured way such
that it can be found and used by anyone with access to the World Wide Web.
Chemical Semantics, Inc. is creating client and server software that allows scientists
to automate their publishing of data into a modern graph database [4], i.e. a Giant
Global Graph (GGG), so that they and others can share their data and use it in a way
that is not now possible with the existing World Wide Web (WWW). Adding
semantics to scientific data and making that data available on the semantic web
makes it possible to do science in a new collaborative way that has the potential to
change science forever [1].
To make the technology of the semantic web available to scientists, one has to
“publish data” in a way that is related to the standard model of journal publishing.
That is, one puts the data (not journal text) into an appropriate form, sends it to a
publisher (of data not journal text), waits for its publication, and then informs
colleagues that they can access the data (not journal article) in a standard fashion
(more and more via the web rather than via hard copy).
The appropriate form for data described above has been clearly defined by the
World Wide Web Consortium (W3C) [5] as the Resource Description Framework
(RDF) standard [6]. This standard is a Graph Database where RDF statements all
take the “triple” form (subject, predicate, and object). For example, (water,
has_boiling_point, 100) is of this triple form. That is, there is a graph arc called
“has_boiling_point” which points from a “water” subject node to a “100” object
node. This form surpasses the normal relational database in its applicability to the
web. Any and all scientific data can be put into this form and data on the semantic
web is stored in servers referred to as triple stores. These triple stores may contain
billions of triples.
A Portal for Quantum Chemistry Data … 5
Chemical Semantics, Inc. operates servers that store scientific data in these triple
stores. Our client software allows scientists to automate the publishing of their data
to these triple stores. Initially, we are focusing on data produced by quantum
chemistry packages although the basic ideas apply to any chemical data including
experimental data. The data produced from these quantum chemistry packages
depends somewhat upon the specific package but for illustrative purposes let’s
assume a “CompChem” package that produces results from ab initio wave function
calculations, such as Self-Consistent-Field (Hartree-Fock SCF) calculations, post
Hartree-Fock correlated calculations, etc. Examples of such packages are Gamess
[7], Gaussian [8], NWChem [9], etc.
The publishing of the results of these calculations ought not to be more difficult
than having a “Publish” button in the GUI or text input of the package. For initial
demonstration, we have implemented such a button in HyperChem [10] as shown in
Fig. 1. The specific use of HyperChem is not relevant and any CompChem [11]
package ought perhaps to have such a button! The founders of Chemical Semantics,
Inc. have a historical tie to Hypercube, Inc. and as such have used HyperChem to
illustrate the basic ideas. What this button in a “CompChem” package does is create
an XML file structure that includes the information about the current molecular
system and any current calculation results resident in the package and sends the data
to the Chemical Semantics, Inc. portal where it is published, given the Authors,
Title, Abstract, Login information and other details that are part of the global setup
prior to hitting the “Publish” button. Our current XML file structure is called CSX
and is related in historical terms to the Chemical Markup Language (CML) file
structure [12]. We use CSX because CML currently does not have certain properties
that we consider not only desirable, but mandatory, such as the ability to deal with
residues as independent fundamental units of biological molecules. Also, CML
does not define many of the quantum chemistry concepts that we consider critical,
We believe CSX encompasses CML as a subset and we could generate a CML file
from CSX easily.
Our Chemical Semantics portal accepts data using a REST or SOAP [13] pro-
tocol and then publishes the data on its servers. The data is available to anyone
around the world with an account at the portal. The details of the portal are
described in another section below.
Our portal allows users to access data at the portal based upon various rules, search
criteria, etc. Semantic web data is usually searched for using a SPARQL Protocol
and RDF Query Language (SPARQL) [14]. Note the recursive acronym.
A SPARQL end point is maintained at the portal which allows queries of various
kinds. SPARQL has features in common with the SQL query language and is
relatively easy to use but requires some experience in forming queries. A natural
language front end would be desirable and many groups are involved in developing
such front ends. A query, for example, could ask how many Density Functional
Theory (DFT) calculations have been done on a specific molecule and with which
functionals and what final total energies. As opposed to querying relational data-
base silos, the query could potentially survey the whole world’s set of such cal-
culations and return with a table of these.
Because the data stored by Chemical Semantics includes a relatively unlimited
number of triples, searching can be very exhaustive while still being very focused.
The data is defined by the ontology so that a search does not return irrelevant
results. Because the data is held by a graph database where a graph node (resource)
uses an arrow to point to another resource, the arrow can simply point to a resource
in a second graph from the first graph and unlike a relational database the data from
different graphs can be merged trivially. This “federation” allows searches to really
use a Giant Global Graph (GGG). As part of the publish activity, data on our portal
can be tagged as private, protected, or public. Any search obviously can return
A Portal for Quantum Chemistry Data … 7
public data. For protected data the author can pass a key to researchers to enable
them to access his/her protected data. An alternative would be to label the publi-
cation private so that only the original authors can access the data.
Data also comes with tags set by the authors that help define the search. For
example an author might add a tag, “used_32bit_gpu” if he/she thought it worth-
while to distinguish calculations on a 32-bit only graphical processing unit from
those on a normal CPU. Once data has been put into a semantic web form elaborate
searches can be performed and the retrieved summary data could also be added as
new data if so desired. Examples of SPAROL queries are shown below after the
SPARQL query language is described.
2 Portal Technologies
The semantic web and our portal use a number of “new” technologies that are
briefly described here. Most of these are new to chemists at this point and the
following therefore provides somewhat of a primer on the semantic web so that the
next section describing the actual operation of publishing and querying using the
facilities of Chemical Semantics, Inc. can be better understood. There are many
good reference books on the semantic web but they generally are written for
computer scientists not chemists.
This tutorial will focus on describing quantum chemistry data as applied to the
semantic web.
The semantic web uses RDF and triple stores to hold data in a graphical database.
The data of Quantum Chemistry needs to be put into this form using an ontology
for Quantum Chemistry. As a practical matter it makes sense to have a way of
structuring computed quantum data prior to converting it into RDF semantic web
data. A fundamental issue here is that the data should always be put into a struc-
tured form from its inception so that computers can be taught this structure. We
have introduced a markup language that we call a Common Standard for eXchange
of quantum data (CSX).
2.1.1 Overview
Chemical Semantics, Inc. uses an Extensible Markup Language (XML) format file
to capture information about quantum chemistry calculations. Specifically, a
Common Standard for eXchange (CSX) file is used to transfer structured data and
8 B. Wang et al.
metadata about calculations to our web portal where it is converted to the Resource
Description Framework (RDF) format appropriate to the semantic web.
While CSX has been developed to allow publication pf quantum chemistry
calculations onto the semantic web, it is a useful standard in and of itself because it
organizes all the important information about a calculation in a format that is
readable by both humans and computers. We suggest that CSX could become a
standard output format for the computational chemistry community and invite
interested readers to contribute to its development.
This section thus describes the current CSX standard for describing data from
quantum chemistry calculations (and possibly related computational or experi-
mental data). The standard specifically includes information describing a publica-
tion, i.e. title, author, etc. because our portal essentially accepts “data publications”
and places them onto the semantic web. CSX is still currently under development
and so please be aware that the description below is dated. A CSX file has a version
number and the current version described here is Version 1.0. Ongoing develop-
ment is aimed at creating Version 2.0 in the near future. For example, aligning our
CSX with JSON for Linked Data (JSON-LD) [15] is appropriate as this new
standard for linked data is closely related to the semantic web.
A CSX file is shown in Fig. 2. The fundamental components are described below.
NameSpaces
The CSX file above includes the basic XML components including a comment that
describes where the CSX file was created. The root element of the XML file is
cs:chemicalSemantics. The attributes at the root include this particular version of
CSX, the local file name for the CSX file, and the relevant namespaces being used.
The principal namespace is cs: which signifies the XML elements defined by
Chemical Semantics, Inc. Other namespaces like xsd: and xsi: are associated with
the XML schema (blueprint) that describes CSX. The namespaces dc: and dcterms:
are part of Dublin Core, a metadata standard that is used for some of the publication
parameters of CSX like title and abstract. The namespace bse: is used to describe
basis sets and is subject to modification as we collaborate with Pacific NW Lab-
oratories on semantic definitions of standard basis sets.
The description given here corresponds to CSX Version 1.0. The definition of CSX
is ongoing. While other sections could be added in future versions of CSX, the
current standard includes three sections:
• molecularPublication
• molecularSystem
• molecularCalculation.
The Chemical Semantics portal essentially accepts “Data Publications”, i.e. data
from computational chemistry computations that are being published at the portal.
Thus, the first section of a CSX file describes the publication itself, including copies
of the input and output files used in the calculation (if available). The second
section describes the molecular system (a set of molecules) that calculations were
performed on. The final section describes the calculation (or calculations) that were
performed on the molecular system.
Authors
A publication can have a number of authors described by their type. The sole author
here is described as cs:corresponding, i.e. the Corresponding Author who is usually
the Principal Investigator (PI) or someone with authority over the publication. The
attribute “type” can also have a value of cs:submitting indicating the author sub-
mitting the publication, or could be empty indicating another author or co-author.
Each author has a name described by cs:creator and an organization described by
cs:organization and an email address described by cs:email.
10 B. Wang et al.
Source
The data being published should have an indication of its source, i.e. the software
package that created the data such as Gamess, NWChem, PSI4, etc. The version of
these software packages should also be indicated. In the example above, the data
came from Release 9 of HyperChem.
For archival purposes, it is also possible to add the text that constitutes the input
file for the calculation as well as the text of the output file. The publication data may
have been extracted by parsing the output file or directly from the software package.
In any event, the input and output file (if present) constitute additional archival data
about the calculation that can be recovered later, if so desired.
Finally, the publication can indicate a set of arbitrary tags that can apply to the
publication and provide an additional way to search for a set of publications. These
tags are independent of searches based on SPARQL and are up to the authors to create.
They may or may not be commonly used depending upon the preference of users.
A set of allowed values for flags is used to characterize publications. These are:
Status
• Preliminary
• Draft
• Final.
A Portal for Quantum Chemistry Data … 11
Visibility
• Private
• Protected
• Public.
Depending upon the desires of the authors, a publication can be set to private
which means only the submitting author with a password can see the publication.
Alternatively, a public publication can be seen by anyone with access to the portal.
Finally, an intermediate visibility is available that is termed “protected”. A pro-
tected publication requires a key (similar to a password) that an author can pass to a
collaborator so that they can see an author’s publication.
Category
System
The molecular system currently has three properties. The traditional charge and
multiplicity of a quantum calculation are two of these. In addition, we define the
system temperature, as well, for statistical mechanical calculations such as
molecular dynamics (MD), Monte Carlo, etc.
Molecule
Each molecule has an id which is the name or other identifier of the molecule.
A default id of “m1”, “m2”, etc. for molecule 1, molecule 2, … is suggested.
A Portal for Quantum Chemistry Data … 13
Atom
All atoms in a molecule must have an id attribute of the form “a1”, “a2”, “a3”, etc.
These identifiers are used to describe the resulting bonds.
An atom has a number of children including the elementSymbol, elementName,
etc.
• elementSymbol—this is just normal symbol such as C, Cl, etc.
• elementName—this is the full name for the element such as Carbon, Chlorine,
etc.
• atomName—for macromolecules an atom may have a pertinent name such as
CA, CB for the alpha, beta carbons in a chain. For normal molecules, the default
names are:
– H
– MainGroup
– Metal
– Row1TM
– Row2TM
– Row3TM
– Lanthanide
– Actinide
– NobleGas
• atomMass—the mass in amu
• formalAtomCharge—integer charge such as +1 for the N in NH4
• calculatedAtomCharge—as used in molecular mechanics coulomb interactions
• x/y/zCoord3D—the three Cartesian coordinates of an atom
• basisSet—the basis set is a property of each atom and may be different for
different atoms
• coordination—describes the other atoms to which this atom is connected.
Bond
Bonds in CSX are a property of atoms. The XML coordination element describes
these, as children of coordination. The bondCount attribute of coordination is the
number of bonds that this atom participates in. Each bond is a child of the atom’s
coordination with attributes id1 and id2 that describe the two connected atoms. This
means that each bond is described twice—as a grandchild of the atom with attribute
14 B. Wang et al.
id1 and as a grandchild of the atom with attribute id2. The content of the XML bond
element describes the bond as single, double, triple, aromatic, or dative. We believe
that having a bond be defined in the context of defining an atom provides better
functionality that defining bonds as an isolated property as per CML.
Attributes
waveFunction
The wave function example in Fig. 6. shows the result for the above 3–21 G
calculation with its orbitals, etc. The far right side of the display is cut off and not all
orbital energies, symmetries, etc. are shown.
16 B. Wang et al.
The wave function has attributes orbitalCount and basisCount. These are gen-
erally the same but some calculations such as those from Gamess may differ here
because of the treatment of 5 spherical or 6 Cartesian d-orbitals.
The child XML elements of the waveFunction include the orbitalEnergies, cs:
orbitalSymmetry, cs:orbitalOccupancy and the orbitals themselves. Each orbital is
identified by its id.
Other Calculations
The RDF standard for a graph database and the associated serialized files, like RDF/
XML or Turtle, contain the scientific data (consisting of URI’s and literals). The
interpretation or meaning of that data, however, requires a vocabulary for defining
the data. That vocabulary is an ontology, in this case an ontology for computational
chemistry. Simple ontologies use a schema for RDF (RDFS) but in general the Web
Ontology Language (OWL) and OWL files are better used to describe the ontology.
An OWL file is usually formatted as RDF/XML for convenience although an OWL
ontology is not to be confused with the fundamental data described by RDF. It is
just simply convenient to use RDF to describe an ontology. This shows the power
of RDF in that the ontology can be described by the same structure as the data.
An OWL file begins by describing the classes (and subclasses) that are basic
entities of the ontology. For example the classes might be MolecularSystem,
Molecule and Atom, where the MolecularSystem is what a calculation is performed
on and which is assumed to be a collection of atoms and molecules. These classes
have ObjectProperties that relate one class to another. The ObjectProperty, has-
Molecule, relates the class MolecularSystem (the Domain) to the class Molecule
(the Range). The ObjectProperty, hasAtom, similarly relates the class Molecule to
the class Atom. The following is a portion of such an ontology expressed in Turtle.
The Gainesville Core ontology is reflected in the prefix gc:
@prefix rdfs: <http://www.w3.org/2000/01/rdf-schema#> .
@prefix owl: <http://www.w3.org/2002/07/owl#> .
@prefix gc: http://purl.org/gc>.
Using such an ontology allows the trivial inference that a MolecularSystem has
atoms although that is not explicitly stated! An ontology also has DataType
Properties that relate a class to a Literal rather than another class. For example, the
18 B. Wang et al.
2.5 SPARQL—Searching
The acronym SPARQL Protocol and RDF Query Language (SPARQL) is officially
recursive but sometimes is referred to as Simple Protocol and RDF Query Lan-
guage. It bears some similarity to SQL but is a W3C standard for querying data on
the semantic web (graph databases) rather than data in relational databases.
There are a number of commercial SPARQL software packages including Vir-
tuoso which Chemical Semantics’ portal uses. The following is an example of a
SPARQL query that searches for all molecules that have 5 atoms and include a
Chlorine atom.
prefix gc: <http://purl.org/gc/>
prefix rdfs: <http://www.w3.org/2000/01/rdf-schema#>
select
?molecule ?moleculeLabel ?inchikey
where {
graph ?graph {
?molecule rdf:type gc:Molecule ;
rdfs:label ?moleculeLabel ;
gc:hasNumberOfAtoms “5”;
gc:hasAtom ?atom;
gc:hasInChIKey ?inchikey .
The query begins with a definition of the namespaces gc: (Gainesville Core defined
by Chemical Semantics, Inc.) and rdfs: (RDF Schema) that it will use. The quantities ?
molecule, ?moleculeLabel, etc. are simply variables with arbitrary names.
The query searches for Molecules that have certain properties such as a label,
number of atoms, etc. The “;” is essentially an “and” and the search ends with a “.”
In addition, to the required Molecule properties, a ?atom found in the molecule
must also be a Chlorine according to the last requirement (Fig. 7),
An example of using the query at the portal of Chemical Semantics, Inc. with a
small demonstration RDF graph results in a table as shown in Fig. 8.
The columns are labelled by the variables used in the query select. It returns here
only the inchiKey of the two molecules but an extension of the query could return
any computed properties of the two molecules that are part of our graph database on
the semantic web.
There are three ways to publish on the portal. The first is to publish directly from
various software packages that produce computational chemistry results. The
number of these will expand as time progresses. For example, versions of PSI 4,
NWChem and HyperChem can publish this way. An example used to develop the
basic ideas is HyperChem, Release 9 with its Publish Button. The other packages
without a GUI require a simple script for publishing.
The second way to publish is to independently create a CSX file that defines the
publication, the molecular system and the calculations and just upload that file to
the portal where it will be translated to a Turtle (*.TTL) file and the data placed
onto the semantic web. One way to do this is to parse an output file with software
such as CSI’s own ChemicalPublisher to create the CSX file as shown in Fig. 9.
Thirdly, one might directly upload to the portal the output file from a compu-
tational package and have that output parsed at the portal where it is put into CSX
form initially and then translated to a TTL file.
The first publication procedure, which uses HyperChem 9 or later, will not only
publish HyperChem calculations but any third-party calculations that HyperChem
has imported such as those from Gamess, Gaussian, Mopac, etc.
For example, HyperChem can parse a Gamess output file. The screen shot in
Fig. 9 shows three Gamess output files containing results for a single point ab initio
SCF calculation, a geometry optimization of structure and a vibrational analysis
A Portal for Quantum Chemistry Data … 21
Fig. 9 Using parser software to parse an output file to create a CSX file
calculation. Once imported into HyperChem 9, these results can be published just as
if they we computed by HyperChem.
While parsing output files is certainly possible, Chemical Semantics, Inc.
expects to work with developers of these computational chemistry packages to help
them install their own “Publish Button”. The Publish Button in HyperChem is
shown in Fig. 1.
In addition to the title of the publication, the authors, their organizations and
e-mail, and the publication abstract, pushing the publication button (which initially
creates a CSX file) adds a number of other things to the publication. The Login
Data… Button allows entering data so that the Publisher Package can use a login ID
and password to directly publish results. The Content Button allows choices to be
made of what is published among the available results as shown in Fig. 10. The
Flags button allows the author to choose to define the current state of the publi-
cation as shown in Fig. 11 or the Visibility (Private, Protected, and Public).
A private publication can be seen only by the authors, a protected publication can
be show to anyone that the authors send a URI to with a key, and a public
publication can be seen by anyone logged into the portal.
It is also possible to add tags (essentially keywords) to any publication as shown
in Fig. 12. These may help in searching. A common set of tags is available as well
as custom tags set by the authors.
22 B. Wang et al.
list of one’s own publications but one can inspect all publications as well depending
upon their visibility (Private, Protected, Public).
One can peruse all publications based upon Author, Title, Category, Tag, etc.
and then view any publication.
One of the problems with existing databases is that the data exists in silos of
isolation. The individual databases are difficult to merge and there is general dif-
ficulty in sharing data because of a lack of universal agreement on the database
schema, column names, etc. A fundamental aspect of the semantic web is its ability
to federate data, i.e. make data available globally. This comes about because of the
global data standards that have been set, because one can merge individual
ontologies easily and because two separate graph databases can be merged just by
adding a single link (predicate) from one graph to another (Fig. 14).
An elementary example of this federation is available at the Chemical Semantics
portal by clicking on the Data Federation Tab. If the molecule for the current
publication is Methyl Chloride, then clicking on the tab brings up something like
that shown in Fig. 15.
This displays the information about Methyl Chloride that exists at the
ChemSpider site of the Royal Society of Chemistry (RSC) and the Chemical
Entities of Biological Interest (ChEBI) site of the European Molecular Biology Lab
(EMBI-EBI).
4 Conclusion
The semantic web offers a new way to publish the data created by Quantum
Chemistry calculations that matches the capabilities of the World Wide Web. As
opposed to isolated silos of data that are difficult to find and/or share, the semantic
web makes sharing of data a fundamental attribute. Our portal is a demonstration of
this new technology and hopefully is a precursor of technology allowing scientist to
finally have a vehicle for the proper sharing of scientific data leading to new and
enhanced capabilities.
References
Frank E. Harris
Abstract We refer to atomic wave functions that contain the interelectron dis-
tances as “explicitly correlated”; we consider here situations in which an explicit
correlation factor rij can occur as a power multiplying an orbital functional form
(a Hylleraas function) and/or in an exponent (producing exponential correlation).
Hylleraas functions in which each wave-function term contains at most one linear
rij factor define a method known as Hylleraas-CI. This paper reviews the analytical
methods available for evaluating matrix elements involving exponentially-correlated
and Hylleraas wave functions; attention is then focused on computation of inte-
grals needed for the kinetic energy. In contrast to orbital-product and exponentially-
correlated wave functions, no general formulas have been developed by others to
relate the kinetic-energy integrals in Hylleraas-CI (or its recent extension by the
Nakatsuji group) to contiguous potential-energy matrix elements. The present paper
provides these missing formulas, obtaining them by using relevant properties of vec-
tor spherical harmonics. Validity of the formulas is confirmed by comparisons with
kinetic-energy integrals obtained in other ways.
1 Introduction
Ever since the first days of quantum mechanics investigators have sought meth-
ods for describing the electronic structures of atoms and molecules that are more
rapidly convergent than superposition-of-configurations (also called configuration-
interaction) expansions of the electronic wave function in orbital products. Probably
It is a pleasure to present this work as part of a tribute to Professor Josef Paldus in celebration
of his eightieth birthday.
F. E. Harris (✉)
Department of Physics, University of Utah, Salt Lake City, UT, USA
e-mail: [email protected]fl.edu
F. E. Harris
Quantum Theory Project, University of Florida, Gainesville, FL, USA
the earliest endeavor of this type was that of Hylleraas, whose study of the He atom
[1] used a wave function that included as a multiplicative factor the explicit appear-
ance of the interelectron distance r12 . Although it was many years before wave func-
tions of this type came into widespread use (probably awaiting the availability of
digital computers), a few landmark studies using such wave functions (which we
identify as traditional Hylleraas functions) were soon carried out, including in par-
ticular a study of the hydrogen molecule by James and Coolidge [2], published in
1936. There followed in 1960 a further study of the hydrogen molecule ground state
by Kolos and Roothaan [3], in 1968 a detailed study of the lithium atom by Larsson
[4], and in 1994 an essentially quantitative computation of the ground state of the
hydrogen molecule by Kolos [5].
Attempts to apply Hylleraas methods to larger systems revealed that the occur-
rence of a wide variety of combinations of rij factors (and higher powers thereof)
led to exceedingly complicated computations, and as early as 1971 it was proposed
by Sims and Hagstrom [6], and independently by Woźnicki [7], to consider con-
figurations (wave function terms) that contained at most a single, linear rij factor.
Methods based on wave functions of this type, now usually referred to as Hylleraas-
CI (Hy-CI), were over time more fully developed and applied to a variety of atomic
problems. Representative work in this area is in [8–11].
An alternative to the Hy-CI development is the use of exponentially-correlated
wave functions. This type of wave function was proposed in the mid-1960s by Bon-
ham [12, 13], but at that time calculations based on it seemed impractical. However,
it was practical to use exponentially-correlated Gaussian orbitals, and work in that
area has been pursued by Rychlewski et al. [14]. Calculations based on exponentially-
correlated Slater-type orbitals finally became practical for small atomic systems after
publication of an extraordinary paper by Fromm and Hill [15].
The possibility of a practical extension to the Hy-CI method (identified by its
proposers as E-Hy-CI) has been examined by the Wang et al. [16], who developed
formulas for the “unlinked” integrals (defined in Sect. 4) that are encountered when
the single rij of a Hy-CI wave function is generalized to a form of the generic type
p
rijij exp(−𝛽ij rij ).
The present contribution reviews some aspects of electronic-structure compu-
tations by these Hylleraas-inspired methods, including a discussion of recently-
discovered methods for simplifying the computation of the kinetic-energy matrix
elements in both Hy-CI and E-Hy-CI.
2 Wave Functions
m m
𝛹 (1, 2) = Yl 1 (𝛺1 )Yl 2 (𝛺2 )𝛷(1, 2) , (1)
1 2
m m m
𝛹 (1, 2, 3) = Yl 1 (𝛺1 )Yl 2 (𝛺2 )Yl 3 (𝛺3 )𝛷(1, 2, 3) , (2)
1 2 3
with
Here Ylm are spherical harmonics, at Condon-Shortley phase [17], and 𝛺i stands for
the angular coordinates of Particle i. Particle i, at position 𝐫i , has radial coordinate ri ,
𝐫ij = 𝐫i − 𝐫j , and rij = |𝐫i − 𝐫j | is the distance between Particles i and j. Traditional
Hylleraas methods use wave functions with all 𝛽 equal to zero; Hylleraas-CI methods
additionally require each wave-function term to have at most one pi nonzero; that pi ,
if present, has the value unity.
3 Exponential Correlation
4 Hylleraas Integrals
Integrals containing various combinations of rij as factors (we call these potential-
energy integrals) will occur in atomic Hylleraas calculations; each integral can be
identified with a diagram that is formed by
(1) Introducing a vertex corresponding to each particle not at the origin of the
coordinate system (usually the nucleus);
(2) For each factor rij or 1∕rij in the integral, drawing a line in the diagram that
connects vertices i and j.
Any vertex not connected to other vertices by a line corresponds to a one-particle
integral that is easily evaluated. Diagrams containing closed loops (with three or
more vertices) are termed linked, and any diagram or part of a diagram that is not
contained in a closed loop is called unlinked. Unlinked integrals and parts can be
evaluated in closed form after introducing the Laplace expansion of each 1∕rij [23]
and/or its generalization to the related quantity rij [24]. An alternative to the Laplace-
type expansion for unlinked integrations is to use a coordinate system in which the
unlinked rij are coordinates. That approach has been followed by Ruiz in [25–28] and
in other papers. Integrations over the particles in closed loops can also be treated
using Laplace-type expansions, but such linked integrations lead to infinite series
that are usually evaluated numerically.
For atomic Hy-CI, the limitations in the occurrence of rij factors cause the
potential-energy integrals to consist only of completely unlinked integrals involv-
ing four or fewer vertices, except for one three-vertex integral containing a linked
product of the form rij rik ∕rjk . This type of integral, often called a “triangle” integral,
was first discussed by Szasz [29].
While the general development for exponentially correlated wave functions in
principle provided closed analytic formulas for the Hylleraas triangle integrals,
those formulas were often more laborious to evaluate than expansions based on
Laplace-type formulas. For triangle integrals with general spherical harmonics, the
most utilized current approach is probably the Levin u-transformation convergence-
acceleration scheme [30] used by Sims and Hagstrom [9].
The diagrams denoting integrals arising in Hy-CI are the same as those occurring
in its extension E-Hy-CI, the only difference being that each diagram line refers to a
p
factor of type rijij exp(−𝛽ij rij ) instead of simply rij .
Matrix Elements for Explicitly-Correlated Atomic Wave Functions 33
5 Kinetic Energy
The indices in Eq. (5) run over all the particles (including any nuclei, whether or not
they are assumed to be of infinite mass), mi is the mass of Particle i, and cos 𝜃ijk is
the cosine of the angle between 𝐫ij and 𝐫ik . It can be evaluated as
Because Eq. (5) is general, its use yields the kinetic energy even when the particles
are all of finite mass, thereby removing the need for an estimate of the nonphysical
quantity called “mass polarization”.
For wave functions containing explicit angular factors (e.g., spherical harmonics),
the kinetic-energy operator requires additional terms. This topic is discussed in [32–
34].
Evaluation of the kinetic energy for exponentially-correlated wave functions was
examined in 1993 by Rebane [35], who showed how the kinetic-energy matrix ele-
ments could be written in terms of suitable potential-energy contributions. Rebane’s
derivation was later simplified by the author’s research group [36]. Unfortunately
Rebane’s formula involves all the exponents of the exponentially-correlated wave
function and does not apply to the usual .
m m
+ 2r12 ∇1 gd (1) ⋅ ∇1 Yl d (1) + 2gd (1)∇1 r12 ⋅ ∇1 Yl d (1) . (10)
d d
r12 + r12
2
− r22
𝐫̂ 1 ⋅ 𝐫̂ 12 = , (15)
2r1 r12
0 = ∇1 gd (1) ⋅ ∇1 Ylm (1) . (16)
Equation (17) corrects a sign error that was present in the corresponding equation of
[37].
One further simplification can now be easily made to the final term of Eq. (17):
m
The orthogonality of 𝐫̂ 1 and ∇1 Yl d (1) permit us to write
d
m 2r2 gd (1) m
2gd (1) 𝐫̂ 12 ⋅ ∇1 Yl d (1) = − 𝐫̂ 2 ⋅ ∇1 Yl d (1) . (18)
d r12 d
𝐱̂ + i𝐲̂ 𝐱̂ − i𝐲̂
𝐞̂ 1 = − √ , 𝐞̂ −1 = √ , 𝐞̂ 0 = 𝐳̂ , (19)
2 2
(23)
]
ld (ld + 1)r12 e−𝛽r12 gd (1)
p
[ p ] m
− + 2∇1 gd (1) ⋅ ∇1 r12 e−𝛽r12 Yl d (1)
r12 d
[ p ]
+ 2r12 e−𝛽r12 ∇1 gd (1) + 2gd (1)∇1 r12 e−𝛽r12 ⋅ ∇1 Yl d (1) ,
p m m
⋅ ∇1 Yl d (1) (27)
d d
which differs from Eq. (10) only by replacement of r12 everywhere it occurs by
r12 e−𝛽r12 . Examination of Eq. (27) shows that we now need to evaluate the new quan-
p
tities
( )
[ ] p(p + 1) 2𝛽(p + 1)
∇21 r12 e−𝛽r12 = + 𝛽 2 r12 e−𝛽r12 ,
p p
2
− (28)
r12 r 12
( )
[ p ] p
∇1 r12 e−𝛽r12 = − 𝛽 r12 e−𝛽r12 𝐫̂ 12 .
p
(29)
r12
Inserting the results from Eqs. (28) and (29) into Eq. (27) and then proceeding as in
Sect. 6, we reach
38 F. E. Harris
⟨ ⟩
ld (ld + 1) − (nd + p)(nd − 1) |f f |
| 12 13 |
K3EHCI= 𝛷abc | 2 | 𝛷def
2 | r |
| 1 |
⟨ ⟩ 2 2
(2nd + p)𝛼d |f f | 𝛼 + 𝛽 ⟨ ⟩
+ 𝛷abc || 12 13 || 𝛷def − d 𝛷abc ||f12 f13 || 𝛷def
2 | 1 |
r 2
⟨ ⟩ ⟨ ⟩
p(nd + p) | | 𝛼d p |f f r 2|
| f12 f13 | | 12 13 1 f12 f13 r2 |
− 𝛷abc | 2 | 𝛷def + 𝛷abc | 2 − |𝛷
2 | r | 2 | r 2 | def
r1 r12
| 12 | | 12 |
⟨ ⟩ ⟨ ⟩
(n − 1)p | 2|
𝛽(nd + 2p + 1) |f f |
| f12 f13 r |
+ d 𝛷abc | 2 2 2 | 𝛷def + 𝛷abc || 12 13 || 𝛷def
2 | r r | 2 | r12 |
| 1 12 |
⟨ ⟩
𝛼d 𝛽 |f f r 2|
| 12 13 12 f12 f13 r1 f12 f13 r2 |
− 𝛷abc | + − |𝛷
2 | r1 r12 r1 r12 || def
|
⟨ ⟩
𝛽(nd − 1) |f f r r22 f12 f13 ||
| 12 13 12
+ 𝛷abc | − 2 | 𝛷def
2 | r2 r1 r12 ||
| 1
√ ( )1∕2 1 ⟨ ⟩
4𝜋 ∑ ld (ld +1)(2ld +1−𝜆) ∑ l + 𝜆 1 ld
+ (−1)me +𝜈 d
3 𝜆=±1 2(2ld + 1) md −𝜈 𝜈 md
𝜈=−1
[ ] [ ⟨ ⟩
∑ l 1 l + 𝜆′ |f f r |
| 12 13 2 | md −𝜈 me +𝜈 mf
× e e
p 𝛷abc | |d e f
me 𝜈 −me −𝜈 | r r2 | ld +𝜆 le +𝜆′ lf
𝜆′ =±1 | 1 12 |
⟨ ⟩]
| f12 f13 r2 | md −𝜈 me +𝜈 mf
−𝛽 𝛷abc | | | .
|d e ′ f
| r1 r12 | ld +𝜆 le +𝜆 lf
(30)
To keep the above formula more compact, we have defined f12 = r12 e−𝛽r12 and f13 =
p
p′ −𝛽 ′ r13
r13 e .
Equation (30) confirms that the kinetic-energy matrix elements in E-Hy-CI reduce
to contiguous potential-energy integrals.
8 Numerical Verification
The formulas for kinetic-energy intergrals in Hy-CI developed here and (in more
detail) in [37] were confirmed by comparing integrals produced using them with
similar integrals computed in other ways by Ruiz [25, 27, 28] and by Sims and
Hagstrom [10]. After making some adjustments needed to achieve consistency (see
[37]), complete agreement with the results of those investigators was obtained.
The errors noted in various equations of [37] arose while transcribing the formu-
las from computer programs and therefore did not affect the numerical verification
process.
Matrix Elements for Explicitly-Correlated Atomic Wave Functions 39
We could not find literature values of E-Hy-CI kinetic-energy integrals for com-
parison with the formula in Sect. 7 of the present contribution and therefore cannot
present data to provide its numerical confirmation.
9 Conclusions
The spherical harmonics Ylm (𝜃, 𝜙), alternatively written Ylm (𝛺), can be defined with
the sign convention chosen by Condon and Shortley [17] (Condon-Shortley phase)
by the Rodrigues formula
(−1)m 2 m∕2 d
l+m
Ylm (𝛺) = Nlm (1 − u ) (u2 − 1)l eim𝜙 , (31)
2l l! dul+m
where u = cos 𝜃 and Nlm is the factor
√
(2l + 1)(l − m)!
Nlm = (32)
4𝜋(l + m)!
m m
Expansion of the spherical harmonic product Yl 1 Yl 2 in the orthonormal set YLM ,
1 2
carried out by taking scalar products with (YLM )∗ , leads after use of Eqs. (33) and
(34) to [ ]
∑ l l L
(−1)M 1 2
m m
Yl 1 (𝛺)Yl 2 (𝛺) = Y M (𝛺). (35)
1 2 m1 m2 −M L
LM
Because harmonics with upper indices m1 and m2 form a product all of whose terms
have the same value of M, Eq. (35) can be simplified by dropping the M summation,
setting M = m1 + m2 .
The Gaunt coefficients can be written in terms of Wigner 3-j symbols [39]. Using
the standard notation for that symbol (an array of l and m values in ordinary paren-
theses), the Gaunt coefficients as defined here can be written
[ ] √ ( )( )
l 1 l2 l3 (2l1 + 1)(2l2 + 1)(2l3 + 1) l1 l2 l3 l1 l2 l3
= . (36)
m1 m2 m3 4𝜋 m1 m2 m3 0 0 0
where the array in angle brackets is our (nonstandard) notation for the Clebsch-
Gordan coefficient. Here all contributing terms must satisfy m1 + m2 = M, so we
can actually reduce Eq. (37) to a single sum over, say, m2 , with m1 set to M − m2 .
The Clebsch-Gordan coefficients can also be written in terms of 3-j symbols:
⟨ ⟩ ( )
j1 j 2 j 3 j1 −j2 +m3
√ j1 j2 j3
= (−1) 2j3 + 1 . (38)
m1 m2 m3 m1 m2 −m3
Well-documented computer programs exist for the evaluation of the 3-j symbols,
making it straightforward to evaluate expressions involving Gaunt or Clebsch-Gordan
coefficients.
References
1 Introduction
bond breaking
A − B ! A∙ + B∙ ,
Here, ZP means all energies (E and Ẽ) are corrected for the zero-point vibrational
effect. The capital letters A and B refer to the corresponding molecular fragments
produced after breaking the bond between A and B, and the small letters a and
b stand for the basis sets attached to the fragments. Before the homolysis reaction
happens, the parent compound optimized within the combined basis set (a ∪ b) has
energy E(ZP)aAB∪ b. After the reaction occurs, A and B radicals optimized each
individually within their own basis sets, a and b, have energies E(ZP)aA and E(ZP)bB,
respectively. To mitigate the inconsistency of the basis sets used before and after
the reaction, the counterpoise correction [10, 11] for the basis-set superposition
error (BSSE) has been adopted in Eq. (1), in which four more single-point energies
with molecular fragments A and B frozen in their geometries within the parent
∪b ∪b
molecule are calculated: Ẽ(ZP)aA(B) with ghost B, Ẽ(ZP)a(A)B with ghost A, Ẽ(ZP)aA,
and Ẽ(ZP)B. For a given ghost structure, the full sets of its basis functions and
b
numerical integration grid points are still present as before, but there are neither
∪b
nuclear charges nor electrons within the ghost. Therefore, both Ẽ(ZP)aA(B) and Ẽ
a∪b
(ZP)(A)B are still calculated within the total combined basis set (a ∪ b).
Equation (1) clearly shows that for every bond breaking occurrence, an accurate
estimate of the BDE involves at least four single-point calculations, two geometry
optimizations, and six vibrational frequency analyses for the ZP correction. Hence,
calculating the BDEs of all bonds of a large molecule to identify the weakest bond
is very labor-intensive and time-consuming. It is thus highly desirable to design a
simpler structural indicator to replace the enormous amount of BDE calculations. In
conventional chemical wisdom, bond length (R), bond order (BO), and bond
energy (BE) are the three commonly used key parameters reflecting the strength of
Effective Bond-Strength Indicators 45
Fig. 1 Relationship between average bond energies (BEa) and bond length (Ra). The straight line
is a least-square linear fit to the data points denoted by blue squares; data points marked by red
circles are excluded from the linear fit. All data are collected in Table 1
On the other hand, many previous studies [4–7] have primarily used the
Mulliken interatomic electron number (MIEN), calculated via the Mulliken
population analysis [13], to identify the trigger bond. Normally, only the bond with
the smallest MIEN from the same type of bonds is selected for further consideration
[4–7]. However, the bond with the smallest MIEN might not necessarily have the
lowest BDE among the bonds within a molecule. Taking 2,4,6-trinitrotoluene
(TNT) for example, the MIENs of all C–NO2 bonds are among the smallest (see
Table 2), indicating one of the C–NO2 bonds to be the trigger bond. However, to
single out which C–NO2 bond to break first, calculations of the BDEs of these
specific C–NO2 bonds have been carried out. In Table 2, the C3–N11 bond has the
smallest MIEN (0.1468), whereas the C1–N14 or C5–N8 bond has the lowest BDE,
237.75 kJ/mol. Clearly, the smallest MIEN may not correspond to the lowest BDE.
Alternatively, within the quantum theory of atoms in molecules [14, 15], Bader
proposed several parameters to indicate the relative bond strength using the idea
associated with bond critical points (BCPs). These BCPs are located at the inter-
atomic surface between a pair of atoms, at which the electron density reaches
minimum in one dimension, yet reaches maximum in the other two dimensions. As
Bader originally suggested, the value of the electron density at such a bond critical
point between a pair of atoms of a chemical bond, ρc, can be used to measure the
strength of the chemical bond. On the other hand, the BE is an integral of the
electron density over the associated interatomic surface between an atomic pair.
Unfortunately, this BE integral has an unknown system-dependent dimensionless
proportionality pre-factor. As a result, Bader’s ρc indicator can only be used to
compare the same kind of bonds between the same pair of atoms within very similar
local chemical environment and might not be able to evaluate the strengths of
different kinds of bonds even within the same molecule.
Effective Bond-Strength Indicators 47
Table 2 (continued)
Molecule Bond Ra WBOb MIENc Wd Me Kf ρgc BDEh
PAM C2–N7 1.477 0.9191 0.1475 0.62 0.10 0.06 1.1958 217.54
C4–N9 1.477 0.9191 0.1475 0.62 0.10 0.06 1.1958 217.54
C6–N10 1.474 0.9241 0.1501 0.63 0.10 0.06 1.2627 262.69
C3–O8 1.340 1.0529 0.3040 0.79 0.23 0.18 1.3156 354.45
O8–C17 1.453 0.8424 0.2252 0.58 0.16 0.09 1.0756 263.82
C17–H20 1.090 0.9267 0.3626 0.85 0.33 0.30 0.9198 398.18
C17–H21 1.090 0.9267 0.3626 0.85 0.33 0.28 0.9198 398.18
C17–H22 1.090 0.9305 0.3856 0.85 0.35 0.28 0.9158 398.18
AMNA N2–N4 1.422 0.9807 0.1603 0.69 0.11 0.08 0.3241 107.87
N2–N3 1.397 1.0575 0.2451 0.76 0.18 0.13 0.3306 248.42
C1–N2 1.463 0.9571 0.2703 0.65 0.18 0.12 0.2575 233.11
C1–H7 1.088 0.9131 0.3737 0.84 0.34 0.29 0.2807 376.14
C1–H8 1.099 0.9240 0.3729 0.84 0.34 0.29 0.2718 376.41
C1–H9 1.090 0.9141 0.3713 0.84 0.34 0.29 0.2789 376.59
N3–H10 1.018 0.8192 0.3215 0.81 0.32 0.25 0.3333 315.18
N3–H11 1.024 0.8231 0.3109 0.80 0.30 0.24 0.3274 314.42
AMNFMC N7=N8 1.244 1.4789 0.2938 1.19 0.24 0.28 0.4313 141.55
N5–N11 1.449 0.9167 0.1381 0.63 0.10 0.06 0.3031 124.16
C1–O3 1.361 0.9845 0.2689 0.72 0.20 0.14 0.3014 326.62
C1–N5 1.414 0.9998 0.1857 0.71 0.13 0.09 0.2963 320.40
O3–C4 1.443 0.8441 0.1903 0.58 0.13 0.08 0.2393 264.75
C4–N7 1.454 0.9896 0.2696 0.68 0.19 0.13 0.2683 267.37
C4–H15 1.093 0.9075 0.3719 0.83 0.34 0.28 0.2853 358.88
N5–C6 1.444 0.9583 0.2199 0.66 0.15 0.10 0.2701 330.94
C6–F10 1.378 0.8574 0.2802 0.62 0.20 0.13 0.2465 418.79
NMDACB C1–N2 1.479 0.9456 0.2486 0.64 0.17 0.11 0.2590 196.21
C1–N3 1.479 0.9685 0.3084 0.66 0.21 0.14 0.2704 170.88
C1–H10 1.098 0.9004 0.3552 0.82 0.32 0.27 0.2774 367.45
N2–N6 1.389 0.9756 0.1878 0.70 0.14 0.10 0.3469 157.49
N3–C5 1.455 1.0026 0.3014 0.69 0.21 0.14 0.2704 294.30
C5–H14 1.106 0.9212 0.3525 0.83 0.32 0.27 0.2666 378.11
a
R is the bond length (in Å)
b
WBO is the Wiberg bond order, defined in Refs. [21, 22]
c
MIEN is the Mulliken interatomic electron number, defined in Ref. [13]
d
W = WBO/R (in Å−1), rounded to the second decimal place for optimal sensitivity
e
M = MIEN/R (in Å−1), rounded to the second decimal place for optimal sensitivity
f
K = W × M (in Å−2), rounded to the second decimal place for optimal sensitivity
g
ρc is the electron density at the bond critical point within Bader’s atoms-in-molecules analysis, defined
in Refs. [14, 15]
h
BDE (in kJ/mol) is the bond dissociation energy calculated according to Eq. (1)
Effective Bond-Strength Indicators 49
H20 H20 H 24
21
H H19 21
H H19 23 H H25
O 7 O O 7 O O 22 O
N N N N N N
O 8 6 14 O O 8 6 16 O O 7 1 19 O
5 5 6
1 1 2
4 2 2 3
H 4 H25 O 5 O
3 3 4
11 N 10 16 N
H 15
11 12
H 26 13
N H N H 27 O N O
O O O O O O
H 22
H20 21 H
19 H H21 20 H
O O H17 17
7 O O8 O
N N N N Cl
7
O 12 1 8 O
16 H 10
5
6 N N
6 1 9
2 O 7
2
3 O
3 2 4
5 O 4 O
24 H 4 3
1 5
N 11 O9 N9 8
N 6
10
H 23 N H22 O H15 O 10
O O N
O O
O F10
17
H
H11 H9 N H16
H8 6 O 6
4
N O 5
2
N 2 N
1 11
H10 3
N2 H7 1 O
N
3 1 3
N N O
4 7
N H9 N O
13 H 8 4
5
H10 N9
O O
H14 H15
H15
Fig. 2 Molecular structures of selected explosive compounds (some hydrogens omitted for
clarity)
50 G.-X. Wang et al.
Hereafter, the five key bond-strength indicators (i.e., R, MIEN, BO, ρc, and
BDE) of all single bonds of these molecules will be carefully compared to seek
better relationships among them. The end goal is to propose simple, sensitive
structural indicators suitable for identifying the trigger bond (or even better, all
weak bonds), thus saving both computational resources and time.
2 Computational Methods
Many studies [1–8, 16, 17], have already shown that the DFT-B3LYP method [18,
19] in combination with the 6-31G* basis set [20] is able to yield accurate ener-
getics, structures, and other molecular properties. In this paper, the same method
was employed to obtain the fully optimized molecular geometries and electronic
structures, including MIEN and Wiberg bond order (WBO) [21, 22], of the chosen
compounds (Fig. 2) within the Gaussian09 program package [23]. To obtain the
values of ρc, the Bader analysis was performed using AIMAll package [24].
Based on Hess’s law [25], conventional counterpoise correction methods [10,
11] were employed to calculate the BDEs according to Eq. (1). Unfortunately, any
existing counterpoise correction methods [10, 11] cannot address the situation when
a bond within a ring is broken. Instead of trying to contemplate suitable ghost atoms
and their linking strategies for ring opening scenarios, we simply supplemented
extra diffuse polarization basis functions to the existing basis sets for the atoms of
the broken bonds to roughly mimic the effects of ghost atoms. Taking the
four-membered ring of NMDACB for example, when breaking the C1–N2 and
C1–N3 bonds, the BSSE was calculated with additional aug-cc-pV5Z Diffuse (1s,
1p, 1d, 1f, 1g, 1h) basis functions [26] separately placed on the C1, N2, and N3
atoms. Numerical tests confirmed that the magnitude of the resulting correction to
the BDEs from this basis-set local enhancement protocol was in line with the
procedure of existing counterpoise correction methods [10, 11].
Four commonly available major bond-strength indicators (i.e., R, MIEN, WBO, and
ρc) are listed along with BDEs in Table 2. For the selected compounds, nearly all
the bonds with the longest R or the smallest WBO do not possess the lowest BDE.
The bond with the lowest BDE is the C–NO2 or N–NO2 bond, but, according to
R and WBO, the weakest bond can draw from all sorts of candidates, e.g., C–C, C–H,
C–NO2, C–Cl, N–H, or O–C bond. Similarly, the smallest values of ρc alone never
correspond to the weakest bonds either. This obviously illustrates that R, WBO, or ρc
cannot be used separately to identify the trigger bond.
Particularly, because of AMNFMC containing the azido group (−N7=N8+=N9−),
when the N7=N8 bond is broken to produce the N2 gas, its BDE is the second lowest
Effective Bond-Strength Indicators 51
among all bonds (merely 141.55 kJ/mol). It is comforting that none of the structural
parameters considered here indicates the N7=N8 bond to be the weakest. However,
this also suggests that whenever exotic bonds (not listed in Table 1) are involved in a
bond breaking process, their BDEs should be calculated to verify the weakest bond
for certainty.
Data in Table 2 also show that MIEN alone can correctly single out the type of
bonds to which the trigger bond belongs. For instance, for the selected compounds,
the easiest bond to break is the C–NO2 or N–NO2 bond, but among which MIEN
sometimes incorrectly picks up the weakest bond. For CDNAPY, the C4–N9 bond
has the smallest MIEN (0.1237) and its BDE is 278.67 kJ/mol, but it is the C2–N8
bond that has the lowest BDE (248.56 kJ/mol) even with a bigger MIEN (0.1335).
Therefore, much sensible structural parameters other than the existing four afore-
mentioned must be proposed for trigger bond identification.
In light of the general fact that WBO and MIEN are related positively to BDE
and inversely to R, W = WBO/R and M = MIEN/R were first conceived to be the
better alternatives. Such an idea comes with no surprise, because the numerator of
the 1/Ra term on the right-hand side of Eq. (2) can be interpreted as WBO = 1 for
single bonds. Then, to take local chemical environment into consideration, Eq. (2)
might suggest a general relationship: BDE/R ∝ WBO/R. (Hereafter, numerical
values of bond-strength indicators will be quoted without units unless otherwise
noted.)
Unfortunately, the data shown in Table 2 again proclaim that the minima of
W and BDE still do not align well. For example, in PAM, the O8–C17 bond has the
smallest W, but the C2–N7 or C4–N9 bond has the lowest BDE. In general, the
results of W closely resemble those of WBO.
Fortunately, the bonds with the smallest M enclose the bond with the lowest
BDE (see Table 2). For TNT, the C1–N14, C3–N11, and C5–N8 bonds all have the
same smallest M value (0.10) and both of the C1–N14 and C5–N8 bonds have the
lowest BDE. For PNT, the C3–N16, C4–N13, and C5–N10 bonds have the smallest
M (0.09) and, among them, the C4–N13 bond is the weakest (BDE = 197.65 kJ/
mol). We then considered the product of W and M, K = W × M, as a new indicator.
According to M, for TNCr, the weakest bonds are the C2–N8 and C6–N12 bonds
(M = 0.11), whereas according to K, the weakest bond of TNCr is solely the C2–N8
bond (K = 0.06) in agreement with its lowest BDE value. Overall, K outperforms
M only marginally (see Fig. 3).
Next, we observed an enlightening fact: ρc can almost always single out the
weakest bond among the weaker bonds of the same type within a molecule iden-
tified by M and K (see Table 2). The only exception might be PNT: among the three
most weak bonds (C3–N16, C4–N13, and C5–N10) sorted out by M and K, ρc chooses
the C5–N10 bond (BDE = 200.63 kJ/mol) to be the weakest whereas the C4–N13
bond has the lowest BDE (197.65 kJ/mol). However, this is not a huge failure
because the difference in the BDEs of the three weak bonds of PNT is less than
3 kJ/mol, well within the error margin of the computational methods employed.
Nonetheless, such promising results strongly advocate that M and K, as two new
trigger bond indicators (TBIs), shall be first used to identify the set of most weak
52 G.-X. Wang et al.
Fig. 3 Relative strengths of all single bonds within TNCr measured by bond-strength indicators:
M, K, and BDE
bonds and ρc can subsequently narrow down the candidates whose BDEs can then
be calculated to pinpoint the trigger bond among them. This can really reduce the
amount of work drastically.
In general, the smaller TBI values almost perfectly map to all weak bonds (with
BDE < 350 kJ/mol) sequentially, and both M and K do indeed capture the weakest
bond once these two TBIs reach minimum (see Table 2). In spite of such a general
success of these two TBIs, neither M nor K can always predict the relative order of
bond strength for strong bonds (with BDE > 350 kJ/mol) within a molecule.
Figure 4 showcases this point succinctly. Such a phenomenon is well expected
because the molecular fragments after breaking a strong bond normally undergo
Fig. 4 Comparison of TBIs (K and M) and BDE of all single bonds within TNCr
Effective Bond-Strength Indicators 53
4 Conclusions
Acknowledgements We are grateful to the grant support from the Natural Sciences and Engi-
neering Research Council (NSERC) of Canada, the National Natural Science Foundation of China
(No. 21403110), and the Natural Science Foundation of Jiangsu Province (No. BK20130755).
This work was mainly carried out at UBC during G.X.W.’s one-year visit to UBC from 16 August
2012 to18 August 2013.
References
1. Xu XJ, Xiao HM, Ju XH, Gong XD, Zhu WH (2006) Computational studies on
polynitrohexaazaadmantanes as potential high energy density materials (HEDMs). J Phys
Chem A 110:5929–5933
2. Qiu L, Xiao HM, Gong XD, Ju XH, Zhu WH (2006) Ab initio and molecular dynamics
studies of crystalline TNAD (trans-1,4,5,8-Tetranitro-1,4,5,8-tetraazadecalin). J Phys Chem A
110:3797–3807
3. Qiu LM, Gong XD, Wang GX, Zheng J, Xiao HM (2009) Looking for high energy density
compounds among 1,3-Bishomopentaprismane derivatives with –CN, –NC, and –ONO2
groups. J Phys Chem A 113:2607–2614
4. Xu XJ, Xiao HM, Gong XD, Ju XH, Chen ZX (2005) Theoretical studies on the vibrational
spectra, thermodynamic properties, detonation properties and pyrolysis mechanisms for
polynitroadamantanes. J Phys Chem A 109:11268–11274
5. Wang GX, Shi CH, Gong XD, Xiao HM (2009) Theoretical investigation on structures,
density, detonation properties and pyrolysis mechanism of the derivatives of HNS. J Phys
Chem A 113:1318–1326
54 G.-X. Wang et al.
6. Wang GX, Gong XD, Yan L, Du HC, Xu XJ, Xiao HM (2010) Theoretical studies on the
structures, density, detonation properties, pyrolysis mechanisms and impact sensitivity of
nitro derivatives of toluenes. J Hazard Mater 177:703–710
7. Wang GX, Gong XD, Du HC, Liu Y, Xiao HM (2011) Theoretical prediction of properties of
aliphatic polynitrates. J Phys Chem A 115:795–804
8. Liu Y, Gong XD, Wang LJ, Wang GX, Xiao HM (2011) Substituent effects on the properties
related to detonation performance and sensitivity for 2,2’,4,4’,6,6’-Hexanitroazobenzene
derivatives. J Phys Chem A 115:1754–1762
9. Jensen F (2007) Introduction to computational chemistry, 2nd edn. Wiley, Chichester, West
Sussex, England
10. Boys SF, Bernardi F (1970) The calculation of small molecular interactions by the differences
of separate total energies. Some procedures with reduced errors. Mol Phys 19:553–566
11. Simon S, Duran M, Dannenberg JJ (1996) How does basis set superposition error change the
potential surfaces for hydrogen bonded dimers? J. Chem. Phys. 105:11024–11031
12. Petrucci RH, Herring FG, Madura JD, Bissonette C (2011) General chemistry, 10th edn.
Pearson Publishing, Toronto, Ontario, Canada
13. Mulliken RS (1962) Criteria for the construction of good self-consistent-field molecular
orbital wave functions, and the significance of LCAO-MO population analysis. J. Chem.
Phys. 36:3428–3439
14. Bader RFW, Tang TH, Tal Y, Biegler-Koenig FW (1982) Properties of atoms and bonds in
hydrocarbon molecules. J Am Chem Soc 104:946–952
15. Bader RFW (1994) Atoms in molecules: a quantum theory. Clarendon Press, Oxford, UK
16. Xiao HM, Xu XJ, Qiu L (2008) Theoretical design of high energy density materials. Science
Press, Beijing, China
17. Wang GX, Gong XD, Liu Y, Du HC, Xu XJ, Xiao HM (2011) Looking for high energy
density compounds applicable for propellant among the derivatives of DPO with −N3,
−ONO2, and −NNO2 groups. J Comput Chem 32:943–952
18. Lee C, Yang W, Parr RG (1988) Development of the Colle-Salvetti correlation-energy
formula into a functional of the electron density. Phys. Rev. B 37:785–789
19. Becke AD (1992) Density-functional thermochemistry, II. The effect of the Perdew-Wang
generalized-gradient correlation correction. J. Chem. Phys. 97:9173–9177
20. Hariharan PC, Pople JA (1973) Influence of polarization functions on MO hydrogenation
energies. Theor. Chim. Acta 28:213–222
21. Wiberg KB (1968) Application of the Pople-Santry-Segal CNDO method to the cyclopropy-
lcarbinyl and cyclobutyl cation and to bicyclobutane. Tetrahedron 24:1083–1096
22. Mayer I (1983) Change bond order and valence in the ab initio SCF theory. Chem Phys Lett
97:270–274
23. Frisch MJ, Trucks GW, Schlegel HB, Scuseria GE, Robb MA, Cheeseman JR, Barone V,
Mennucci B, Cossi M, Scalmani G, Petersson GA, et al (2009) Gaussian 09, revision A. 02,
Gaussian, Inc., Wallingford, CT
24. Bader RFW Atoms in molecules, McMaster University, Hamilton, Ontario, Canada. (http://
www.chemistry.mcmaster.ca/bader/aim/ Accessed 3 Jan 2016)
25. Engel T, Reid P, Hehre W (2013) Physical chemistry, 3rd edn. Boston, USA, Pearson
26. Kendall RA, Dunning TH, Harrison RJ (1992) Electron affinities of the first-row atoms
revisited. Systematic basis sets and wave functions. J. Chem. Phys. 96:6796–6806
Advanced Relativistic Energy Approach
in Electron-Collisional Spectroscopy
of Multicharged Ions in Plasmas
1 Introduction
An accurate data about spectra, radiative decay widths and probabilities, oscillator
strengths, electron-collision strengths, collisional excitation and de-excitation rates
for atoms and especially ions are of a great interest for different applications,
namely, astrophysical analysis, laboratory, thermonuclear plasmas diagnostics,
fusion research, laser physics etc. [1–30]. It is also very important for studying
energy, spectral and radiative characteristics of a laser-produced hot and dense
plasmas [1, 2, 9, 10]. Above other important factors to studying electron-collisional
spectroscopy of ions one should mention the X-ray laser problem. It has stimulated
a great number of papers, devoted to modelling the elementary processes in laser,
collisionally pumped plasmas (see [3, 4] and Refs. therein) and construction of the
first VUV and X-ray lasers with using plasmas of Li-, Ne-like ions as an active
medium. Very useful data on the X-lasers problem are collected in the papers by
Ivanova et al. (see [3–6] and Refs. therein). From the other side, studying spectra of
ions in plasmas remains very actual in order to understand the plasmas processes
themselves. In most plasmas environments the properties are determined by the
electrons and the ions, and the interactions between them. The electron-ion colli-
sions play a major role in the energy balance of plasmas. For this reason, modelers
and diagnosticians require absolute cross sections for these processes. The cross
sections for electron-impact excitation of ions are needed to interpret spectroscopic
measurements and for simulations of plasmas using collisional-radiative models. At
present time a considerable interest has been encapsulated to studying elementary
atomic processes in plasmas environments (for example, see [1–30] and Refs.
therein) because of the plasmas screening effect on the plasmas-embedded atomic
systems. In many papers the calculations of various atomic and ionic systems
embedded in the Debye plasmas have been performed [11, 12, 16, 29, 30]. Cal-
culation of emission spectra of the plasmas ions based in the precise theoretical
techniques is practical tool, which may be used instead of very expensive sophis-
ticated experiments. Nevertheless, there are known principal theoretical problems to
be solved in order to receive the correct description of master parameters of the
elementary atomic processes in laser, collisionally pumped plasmas. First of all,
speech is about development of the advanced quantum-mechanical models for the
further accurate computing oscillator strengths, electron-collisional strengths and
rate coefficients for atomic ions in plasmas, including the Debye plasmas. As
usually, a correct accounting of the relativistic, exchange-correlation, a plasmas
environment effects is of a great importance. To say strictly, solving of the whole
problem requires a development of the quantum-electrodynamical approach as the
most consistent one to problem of the Coulomb many-body system.
In this chapter we present the fundamentals of an advanced relativistic energy
approach, based on the Gell-Mann and Low formalism, to studying spectroscopic
characteristics of the multicharged ions in the Debye plasmas, in particular, com-
puting the electron-ion collision strengths, cross-sections etc. The approach is
combined with relativistic many-body perturbation theory (PT) with the Debye
shielding model Hamiltonian for electron-nuclear and electron-electron systems.
The optimized one-electron representation in the PT zeroth approximation is con-
structed by means of the correct treating the gauge dependent multielectron con-
tribution of the lowest PT corrections to the radiation widths of atomic levels. It is
worth to remind that the method of the relativistic many-body PT formalism is
constructed on the base of the same ideas as the well-known PT approach with the
model potential zeroth approximation by Ivanov-Ivanova et al. [31–44]. However
there are a few fundamental differences. For example, in our case the PT zeroth
approximation [51, 54] is in fact the Dirac- Debye-Hückel one. In order to calculate
the radiative and collisional parameters an effective gauge-invariant version of
Advanced Relativistic Energy Approach in Electron-Collisional … 57
Let us start our consideration from formulation relativistic many-body PT with the
Debye shielding model Dirac Hamiltonian for electron-nuclear and
electron-electron systems. Formally, a multielectron atomic systems (multielectron
atom or multicharged ion) is described by the relativistic Dirac Hamiltonian (the
atomic units are used) as follows:
H = ∑ hðri Þ + ∑ V ri rj . ð1Þ
i i>j
1 − αi αj
V ri rj = exp iωij rij ⋅ , ð2Þ
rij
where ωij is the transition frequency; αi,αj are the Dirac matrices.
In order to take into account the plasmas environment effects already in the PT
zeroth approximation we use the known Yukawa-type potential of the following
form:
where ra, rb represent respectively the spatial coordinates of particles, say, A and B
and Za, Zb denote their charges.
The potential (3) is (look, for example, [23–28] and Refs therein) well known,
for example, in the classical Debye-Hückel theory of plasmas. The plasmas envi-
ronment effect is modelled by the shielding parameter μ, which describes a shape of
the long-rang potential. The parameter μ is connected with the plasmas parameters
such as temperature T and the charge density n as follows:
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
μ∼ e2 n ̸ kB T . ð4aÞ
Here e is the electron charge and kB is the Boltzman constant. The density n is
given as a sum of the electron density Ne and the ion density Nk of the k-th ion
species with the nuclear charge qk:
n = Ne + ∑ q2k Nk . ð4bÞ
k
It is very useful to remind the simple estimates for the shielding parameter. For
example, under typical laser plasmas conditions of T ∼ 1 keV and n ∼ 1023 cm−3
the parameter μ is of the order of 0.1 in atomic units. By introducing the
Yukawa-type electron-nuclear attraction and electron-electron repulsion potentials,
the Dirac-Debye shielding model Hamiltonian for electron-nuclear and
electron-electron subsystems is given in atomic units as follows [28]:
1 − αi αj
H = ∑ ½αcp − βmc − Z expð − μri Þ ̸ri + ∑
2
expð − μrij Þ, ð5Þ
i i>j rij
where c is the velocity of light and Z is a charge of the atomic ion nucleus.
The formalism of the relativistic many-body PT is further constructed in the
same way as the PT formalism in Refs. [31–44]. In the PT zeroth approximation
one should use a mean-field potential, which includes the Yukawa-type potential
(insist of the pure Coulomb one) plus exchange Kohn-Sham potential and addi-
tionally the modified Lundqvist-Gunnarsson correlation potential (with the opti-
mization parameter b) as in Refs. [28–30, 49, 50]. As alternative one could use an
optimized model potential by Ivanova-Ivanov (for Ne-like ions) [31], which is
Advanced Relativistic Energy Approach in Electron-Collisional … 59
calibrated within the special ab initio procedure within the relativistic energy
approach [37, 38].
The most complicated problem of the relativistic PT computing the radiative and
collisional characteristics of the multielectron atomic systems is in an accurate,
precise accounting for the exchange-correlation effects (including polarization and
screening effects, a continuum pressure etc.) as the effects of the PT second and
higher orders. Using the standard Feynman diagram technique one should consider
two kinds of diagrams (the polarization and ladder ones), which describe the
polarization and screening exchange-correlation effects. The polarization diagrams
take into account the quasiparticle (external electrons or vacancies) interaction
through the polarizable core, and the ladder diagrams account for the immediate
quasiparticle interaction. The detailed description of the polarization diagrams and
the corresponding analytical expressions for matrix elements of the polarization
quasiparticles interaction (through the polarizable core) potential are presented in
Refs. [36, 45–50]. An effective approach to accounting for the polarization dia-
grams contributions is in adding the effective two- quasiparticle polarizable operator
into the PT first order matrix elements. In Ref. [36] the corresponding
non-relativistic polarization functional has been derived. More correct relativistic
expression has been presented in the Refs. [45, 46, 49].
Here Φo is the state of the ion with the closed shells (ground state of the Ne-like
ion); Ji is the total angular moment of the initial target state; indices iv, i.e. are
related to the initial states of a vacancy and an electron; indices εin and εsc are the
incident and scattered energies, respectively to the incident and scattered electrons.
The initial state of the system “atom plus free electron” can be written as
where Φo is the state of an ion with closed electron shells (ground state of Ne-like
ion), |I> represents three-quasiparticle (3QP) state, and |F> represents the
one-quasiparticle (1QP) state.
The justification of the energy approach in the scattering problem is in details
described in Refs. [38, 41–44, 49]. The scattered part of energy shift Im ΔE appears
firstly in the atomic PT second order (the fourth order of the QED PT) in the form
of integral over the scattered electron energy εsc:
Z
dεsc Gðεiv , εie , εin , εsc Þ ̸ðεsc − εiv − εie − εin − i0Þ ð7Þ
σ = − 2 ImΔE ð8bÞ
− Yuk
where QCoul
λ + QBr
λ is the sum of the Coulomb-Yukawa and Breit matrix ele-
− Yuk
ments. The Coulomb part QCoul
λ contains the radial Rλ and angular Sλ integrals
as follows:
62 A. V. Glushkov et al.
− Yuk
QCoul
λ = fRλ ð1243ÞSλ ð1243Þ + Rλ ð1̃243̃ÞSλ ð1̃243̃Þ
ð11Þ
+ Rλ ð12̃4̃3ÞSλ ð12̃4̃3Þ + Rλ ð1̃2̃4̃3ÞS
̃ λ ð1̃2̃4̃3Þg.
̃
Here the tilde designates that the large radial Dirac component f must be
replaced by the small Dirac component g, and instead of li , lĩ = li − 1 should be
taken for ji < li and lĩ = li + 1 for ji > li .
The Breit part can be expressed as follows [39]:
QBr Br Br Br
λ = Qλ, λ − 1 + Qλ, λ + Qλ, λ + 1 ð12Þ
λ = Rλ ð124̃3ÞS
QBr ̃ lλ ð124̃3Þ̃ + Rλ ð1̃2̃43ÞSlλ ð1243Þ +
ð13Þ
+ Rl ð1̃24̃3ÞSlλ ð1̃24̃3Þ + Rl ð12̃43̃ÞSlλ ð12̃43̃Þ .
The details of their computing can be found in Refs. [36–54]. The modified PC
code ‘Superatom-ISAN” (version-93) has been used in all calculations.
Here we present the results of computing the radiative and collisional characteristics
(energy shifts, oscillator strengths, electron-ion cross-sections and collision
strengths) for the Be-, Ne-like ions of Ar, Ni and Kr (Z = 18–36) embedded to the
plasmas environment. Let us remind (see Refs. [11, 12, 16, 28, 39]) that the Be- and
Ne-like ions play an important role in the diagnostics of a wide variety of labo-
ratory, astrophysical, thermonuclear plasmas. Firstly, we list our results on energy
shifts and oscillator strengths for transitions 2s2-2s1/22p1/2,3/2 in spectra of the
Be-like Ni and Kr. The plasmas parameters are as follows: ne = 1022−1024cm−3,
T = 0.5–2 keV (i.e. μ ∼ 0.01–0.3). In Tables 1 and 2 we list the results of
Table 1 Energy shifts ΔE (cm−1) for the 2 s2-[2s1/22p3/2]1 transition in spectra of the Be-like Ni
and Kr ions for different values of the ne (cm−3) and T (in eV) (see explanations in text)
ne 1022 1023 1024 1022 1023 1024
Z kT Li et al. Li et al. Li et al. Our data Our data Our data
NiXXV 500 31.3 292.8 2639.6 33.8 300.4 2655.4
1000 23.4 221.6 2030.6 25.7 229.1 2046.1
2000 18.0 172.0 1597.1 20.1 179.8 1612.5
I-S 8.3 86.6 870.9
KrXXXIII 500 21.3 197.9 2191.9 27.2 215.4 2236.4
1000 15.5 150.5 1659.6 21.3 169.1 1705.1
2000 11.5 113.5 1268.0 16.9 128.3 1303.8
Advanced Relativistic Energy Approach in Electron-Collisional … 63
Table 2 Oscillator strengths gf for the 2 s2-[2s1/22p3/2]1 transition in spectra of the Be-like ion of
Ni for different values of the ne (cm−3) and T (in eV) (gf0–the gf value for free ion)
ne 1022 1023 1024 1022 1023 1024
kT gf0 gf gf gf gfo:our gf: our gf: our gf: our
Li et al. Li et al. Li et al. Li et al. data data data data
500 0.1477 0.1477 0.1478 0.1487 0.1480 0.1480 0.1483 0.1495
1000 0.1477 0.1477 0.1482 0.1480 0.1483 0.1495
2000 0.1477 0.1477 0.1481 0.1479 0.1482 0.1493
I-S 0.1477 0.1477 0.1479
Table 3 The electron-collision strengths for Ne-like Ar excitation from the ground state for
impact electron energy 0.75 keV (numbers in brackets denote the multiplicative powers of ten)
Transition Level J [41] [29] Present data
01–02 2s 2p 0 1,303[−03] 1,415[−03] 1,498[−03]
3 2p3/23s1/2 1 9,017[−03] 9,224[−03] 9,286[−03]
4 2p1/23s1/2 0 2,587[−04] 2,724[−04] 2,783[−04]
5 2p1/23s1/2 1 2,241[−02] 2,342[−02] 2,394[−02]
6 2p3/23p3/2 1 3,456[−03] 3,635[−03] 3,699[−03]
7 2p3/23p3/2 3 2,911[−03] 2,998[−03] 3,065[−03]
8 2p3/23p1/2 2 4,795[−03] 4,922[−03] 4,988[−03]
9 2p3/23p1/2 1 1,033[−03] 1,213[−03] 1,254[−03]
10 2p3/23p3/2 2 6,451[−03] 6,535[−03] 6,597[−03]
11 2p1/23p1/2 1 9,641[−04] 9,993[−04] 1,088[−03]
12 2p1/23p1/2 0 8,794[−04] 8,927[−04] 8,992[−04]
13 2p1/23p3/2 2 7,814[−03] 7,978[−03] 8,113[−03]
14 2p1/23p3/2 1 8,561[−04] 8,723[−04] 9,005[−04]
15 2p3/23p3/2 0 8,670[−02] 8,735[−02] 8,802[−02]
16 2p3/23d3/2 0 1,136[−03] 1,244[−03] 1,296[−03]
17 2p3/23d3/2 1 4,129[−03] 4,327[−03] 4,389[−03]
18 2p3/23d5/2 2 5,227[−03] 5,546[−03] 5,601[−03]
19 2p3/23d5/2 4 3,512[−03] 3,678[−03] 3,714[−03]
20 2p3/23d3/2 3 3,994[−03] 4,133[−03] 4,185[−03]
generation of laser radiation in the short-wave spectral range [3, 4]. Besides, it is
obviously more complicated case in comparison with previous one. Here an
accurate account of the excited, Rydberg, autoionization and continuum states can
play a critical role.
In Table 4 we present the theoretical values of the collisional excitation rates
(CER) and collisional de-excitation rates (CDR) for Ne-like argon transition
between the Rydberg states and from the Rydberg states to the continuum states with
Table 4 The collisional excitation (CER) and de-excitation (CDR) rates (in cm3/s) for Ne-like
argon in plasmas with parameters: ne = 1019−20 cm−3 and electron temperature Te = 20 eV
Parameters ne, cm−3 RMPPT RMPPT RMPPT Present Present Present
results results results
Transition 1 → 2 1 → 3 2 → 3 1 → 2 1 → 3 2 → 3
CDR 1.0 + 19 5.35 − 10 1.64–10 1.13 − 09 5.77 − 10 1.92 − 10 1.28 − 09
(i → i; k)
1.0 + 20 5.51 − 10 1.60 − 10 1.12 − 09 5.94 − 10 1.78 − 10 1.25 − 09
Transition 2 → 1 3 → 1 3 → 2 2 → 1 3 → 1 3 → 2
CER 1.0 + 19 5.43 − 10 5.39 − 12 2.26 − 11 5.79 − 10 1.88 − 12 2.64 − 11
(i → i; k)
1.0 + 20 3.70 − 10 8.32 − 12 2.30 − 11 4.85 − 10 1.13 − 11 2.78 − 11
Advanced Relativistic Energy Approach in Electron-Collisional … 65
Table 5 The collisional excitation (CER) and de-excitation (CDR) rates (in cm3/s) for Ne-like
argon in plasmas with parameters: ne = 1019−20 cm−3 and electron temperature Te = 40 eV (our
data)
Parameters ne, cm−3 Present results Present results Present results
Transition 1 → 2 1 → 3 2 → 3
CDR (i → i;k) 1.0 + 19 3.18 − 10 8.45 − 11 6.81 − 10
1.0 + 20 5.02 − 10 1.56 − 10 4.99 − 10
Transition 2 → 1 3 → 1 3 → 2
CER (i → i;k) 1.0 + 19 5.33 − 10 5.63 − 10 7.11 − 11
1.0 + 20 7.67 − 10 6.94 − 11 8.93 − 11
66 A. V. Glushkov et al.
Acknowledgements The authors are very much thankful to Prof. Jean Maruani and Prof. Alex
Wang for invitation to make contributions on the QSCP-XXI workshop (Vancouver, Canada). The
useful comments of the anonymous referees are very much acknowledged too.
References
1. Oks E (2010) In: Oks E, Dalimier E, Stamm R, Stehle C, Gonzalez MA (eds) Spectral line
shapes in plasmas and gases. Int J Spectr 1:852581
2. Griem HR (1974) Spectral line broadening by plasmas. Academic Press, New York
3. Ivanova EP (2011) Phys Rev A 84:043829
4. Ivanova EP, Grant IP (1998) J Phys B Mol Opt Phys 31:2871; Ivanova EP, Zinoviev NA
(2001) Phys Lett A 274:239
5. Khetselius OYu (2011) Quantum structure of electroweak interaction in heavy finite
fermi-systems. Astroprint, Odessa
6. Glushkov AV (1991) Opt Spectrosc 70:555
7. Malinovskaya SV, Glushkov AV, Khetselius OYu, Svinarenko AA, Mischenko EV,
Florko TA (2009) Int J Quant Chem 109(14):3325
8. Glushkov AV, Loboda AV, Gurnitskaya EP, Svinarenko AA (2009) Phys Scripta
T135:014022
9. Glushkov AV, Khetselius OYu, Svinarenko AA (2013) Phys Scripta T153:014029
10. Svinarenko AA (2014) J Phys Conf Ser 548:012039
Advanced Relativistic Energy Approach in Electron-Collisional … 67
11. Paul S, Ho YK (2010) J Phys B At Mol Opt Phys 43:065701; Kar S, Ho YK (2011) J Phys B
At Mol Opt Phys 44:015001
12. Glushkov AV (2005) Atom in electromagnetic field. KNT, Kiev
13. Yongqiang L, Jianhua W, Yong H, Jianmin Y (2008) J Phys B At Mol Opt Phys 41:145002
14. Glushkov AV, Mansarliysky VF, Khetselius OYu, Ignatenko AV, Smirnov AV, Prepelitsa GP
(2017) J Phys Conf Ser 810:012034
15. Glushkov AV, Loboda AV (2007) J Appl Spectr 74:305, Springer
16. Saha B, Fritzsche S (2007) J Phys B At Mol Opt Phys 40:259
17. Marrs R, Levine M, Knapp D, Henderson J (1988) Phys Rev Lett 60:1715
18. Zhang H, Sampson D, Clark R, Mann J (1988) Atom Dat Nuc Dat Tabl 37:17
19. Reed KJ (1988) Phys Rev A 37:1791; Talukder MR (2008) Appl Phys Lasers Opt 93:576
20. Smith ACH, Bannister ME, Chung YS, Djuric N, Dunn GH, Wallbank B, Woitke O (1999)
Phys Scripta T80:283
21. Buyadzhi VV (2015) Photoelectronics 24:128
22. Zeng S, Liu L, Wang JG, Janev RK (2008) J Phys B At Mol Opt Phys 41:135202
23. Okutsu H, Sako T, Yamanouchi K, Diercksen GHF (2005) J Phys B At Mol Opt Phys 38:917
24. Nakamura N, Kavanagh AP, Watanabe H, Sakaue HA, Li Y, Kato D, Curell FJ, Ohtani S
(2007) J Phys Conf Ser 88:012066
25. Bannister ME, Djuric N, Woitke O, Dunn GH, Chung Y-S, Smith ACH, Wallbank B,
Berrington KA (1999) Int J Mass Spectr 192:39
26. Badnell NR (2007) J Phys Conf Ser 88:012070; Griffin DC, Balance CP, Mitnik DM,
Berengut JC (2008) J Phys B At Mol Opt Phys 41:215201
27. Glushkov AV, Malinovskaya SV, Prepelitsa G, Ignatenko V (2005) J Phys Conf Ser 11:199
28. Malinovskaya SV, Glushkov AV, Khetselius OYu, Loboda AV, Lopatkin YuM,
Svinarenko AA, Nikola LV, Perelygina TB (2011) Int J Quant Chem 111:288
29. Glushkov AV, Khetselius OYu, Loboda AV, Ignatenko AV, Svinarenko AA,
Korchevsky DA, Lovett L (2008) Spectr Line Shapes. AIP Conf Pro 1058:175
30. Buyadzhi VV, Chernyakova YuG, Antoshkina OA, Tkach TB (2017) Photoelectronics 26:94
31. Ivanov LN, Ivanova EP (1979) Atom Dat Nucl Dat Tabl 24:95
32. Driker MN, Ivanova EP, Ivanov LN, Shestakov AF (1982) J Quant Spectr Rad Transfer
28:531
33. Ivanov LN, Letokhov VS (1985) Com Mod Phys D 4:169
34. Vidolova-Angelova E, Ivanov LN, Ivanova EP, Angelov DA (1986) J Phys B At Mol Opt
Phys 19:2053
35. Ivanov LN, Ivanova EP, Aglitsky EV (1988) Phys Rep 166:315
36. Ivanova EP, Ivanov LN, Glushkov AV, Kramida AE (1985) Phys Scripta 32:513
37. Glushkov AV, Ivanov LN (1992) Phys Lett A 170:33
38. Glushkov AV, Ivanov LN, Ivanova EP (1986) Autoionization phenomena in atoms. Moscow
University Press, Moscow, pp 58–160
39. Ivanova EP, Glushkov AV (1986) J Quant Spectr Rad Transfer 36:127
40. Ivanova EP, Gulov AV (1991) Atom Dat Nuc Dat Tabl 49:1
41. Ivanov LN, Ivanova EP, Knight L (1993) Phys Rev A 48:4365
42. Ivanov LN, Ivanova EP, Knight L, Molchanov AG (1996) Phys Scripta 53:653
43. Glushkov AV, Ivanov LN (1992) Preprint of ISAN. AS N-1 Moscow-Troitsk
44. Glushkov AV, Ivanov LN (1993) J Phys B At Mol Opt Phys 26:L379
45. Glushkov AV (1990) Sov Phys J 33(1):1
46. Glushkov AV (1990) J Str Chem 31(4):529
47. Glushkov AV (1992) JETP Lett 55:97
48. Glushkov A V (2012) Quantum systems in chemistry and physics. In: Nishikawa K,
Maruani J, Brändas E, Delgado-Barrio G, Piecuch P (eds) Progress in theoretical chemistry
and physics, vol 26. Springer, Dordrecht, pp 231–252
49. Glushkov AV (2008) Relativistic quantum theory. Quantum mechanics of atomic systems.
Astroprint, Odessa, p 700
50. Khestelius OYu (2008) Hyperfine structure of atomic spectra. Astroprint, Odessa, p 210
68 A. V. Glushkov et al.
51. Glushkov AV (2013) Advances in quantum methods and applications in chemistry, physics
and biology. In: Hotokka M, Maruani J, Brändas J, Delgado-Barrio G (eds) Progress in
theoretical chemistry and physics, vol 27. Springer, Cham, pp 161–177
52. Glushkov AV, Khetselius OYu, Svinarenko AA, Prepelitsa GP (2010) In: Duarte FJ
(ed) Coherence and ultrashort pulsed emission, InTech, Rijeka, pp 159–186
53. Svinarenko AA, Glushkov AV, Khetselius OYu, Ternovsky VB, Dubrovskaya YuV,
Kuznetsova AA, Buyadzhi VV (2017) In: Orjuela JEA (ed) Rare earth element, InTech,
pp 83–104
54. Glushkov AV, Khetselius OYu, Svinarenko AA, Buyadzhi VV, Ternovsky VB,
Kuznetsova AA, Bashkarev PG (2017) In: Uzunov DI (ed) Recent studies in perturbation
theory, InTech, pp 131–150
55. Buyadzhi VV, Glushkov AV, Lovett L (2014) Photoelectronics 23:38
56. Buyadzhi VV, Glushkov AV, Mansarliysky VF, Ignatenko AV, Svinarenko AA (2015)
Sensor Electr Microsyst Techn 12(4):27
57. Glushkov AV, Khetselius OYu, Malinovskaya SV (2008) Frontiers in quantum systems in
chemistry and physics. In: Wilson S, Grout PJ, Maruani J, Delgado-Barrio G, Piecuch P
(eds) Progress in theoretical chemistry and physics, vol 18. Springer, Dordrecht, pp 525–541
58. Glushkov AV, Khetselius OYu, Malinovskaya SV (2008) Europ Phys J ST 160:195
59. Glushkov AV, Khetselius OYu, Malinovskaya SV (2008) Mol Phys 106:1257
60. Glushkov AV, Khetselius OY, Svinarenko AA (2012) Advances in the theory of quantum
systems in chemistry and physics. In: Hoggan P, Maruani J, Brändas E, Delgado-Barrio G,
Piecuch P (eds) Progress in theoretical chemistry and physics, vol 22. Springer, Dordrecht,
pp 51–68
61. Glushkov AV, Khetselius OYu, Lovett L (2010) Advances in the theory of atomic and
molecular systems. In: Piecuch P, Maruani J, Delgado-Barrio G,Wilson S (eds) Progress in
theoretical chemistry and physics, vol 20. Springer, Dordrecht, pp 125–152
62. Glushkov AV, Khetselius YO, Loboda AV, Svinarenko AA (2008) Frontiers in quantum
systems in chemistry and physics. In: Wilson S, Grout PJ, Maruani J, Delgado-Barrio G,
Piecuch P (eds) Progress in theoretical chemistry and physics, vol 18. Springer, Dordrecht, pp
543–560
63. Glushkov AV, Svinarenko AA, Khetselius OYu, Buyadzhi VV, Florko TA, Shakhman AN
(2015) Frontiers in quantum methods and applications in chemistry and physics. In:
Nascimento M, Maruani J, Brändas E, Delgado-Barrio G (eds) Progress in theoretical
chemistry and physics, vol 29. Springer, Cham, pp 197–217
64. Khetselius OYu, Zaichko PA, Smirnov AV, Buyadzhi VV, Ternovsky VB, Florko TA,
Mansarliysky VF (2017) Quantum systems in physics, chemistry, and biology. In: Tadjer A,
Pavlov R, Maruani J, Brändas E, Delgado-Barrio G (eds) Progress in theoretical chemistry
and physics, Springer, Cham, pp 271–281
65. Glushkov AV, Buyadzhi VV, Kvasikova AS, Ignatenko AV, Kuznetsova AA, Prepelitsa GP,
Ternovsky V B (2017) Quantum systems in physics, chemistry, and biology. In: Tadjer A,
Pavlov R, Maruani J, Brändas E, Delgado-Barrio G (eds) Progress in theoretical chemistry
and physics, vol 30. Springer, Cham, pp 169–180
66. Glushkov AV, Rusov VD, Ambrosov SV, Loboda AV (2003) In: Fazio G, Hanappe F
(eds) New projects and new lines of research in nuclear physics. World Scientific, Singapore,
pp 126–132
67. Glushkov O. Khetselius E Gurnitskaya, Loboda A, Florko VD, Sukharev Lovett L (2008)
Frontiers in quantum systems in chemistry and physics. In: Wilson S, Grout PJ, Maruani J,
Delgado-Barrio G, Piecuch P (eds) Progress in theoretical chemistry and physics, vol 18.
Springer, Dordrecht, pp 507–524
68. Glushkov AV, Ambrosov SV, Loboda AV, Gurnitskaya EP, Khetselius OY (2006) Recent
advances in the theory of chemical and physical systems. In: Julien P, Maruani J, Mayou D,
Wilson S, Delgado-Barrio G (eds). Progress in theoretical chemistry and physics, vol 15.
Springer, Dordrecht, pp 285–299
Advanced Relativistic Energy Approach in Electron-Collisional … 69
69. Glushkov AV, Ambrosov SV, Loboda AV, Chernyakova YuG, Svinarenko AA, Khetselius
OYu (2004) Nucl Phys A Nucl Hadr Phys 734:21
70. Glushkov AV, Malinovskaya SV, Loboda AV, Shpinareva IM, Prepelitsa GP (2006) J Phys
Conf Ser 35:420
71. Glushkov AV, Ambrosov SV, Loboda AV, Gurnitskaya EP, Prepelitsa GP (2005) Int J Quant
Chem 104:562
72. Glushkov A, Malinovskaya S, Loboda A, Shpinareva I, Gurnitskaya E, Korchevsky D (2005)
J Phys Conf Ser 11:188
73. Glushkov AV, Malinovskaya SV, Chernyakova YG, Svinarenko AA (2004) Int J Quant
Chem 99:889
74. Glushkov AV, Ambrosov SV, Ignatenko AV, Korchevsky DA (2004) Int J Quant Chem
99:936
75. Glushkov AV, Malinovskaya SV, Sukharev DE, Khetselius OYu, Loboda AV, Lovett L
(2009) Int J Quant Chem 109:1717
76. Khetselius OYu (2009) Int J Quant Chem 109:3330
77. Khetselius OYu (2009) Phys Scripta T135:014023
78. Khetselius OYu (2012) Quantum systems in chemistry and physics. In: Nishikawa K,
Maruani J, Brändas E, Delgado-Barrio E, Piecuch P (eds) Progress in theoretical chemistry
and physics, vol 26. Springer, Dordrecht, pp 217–229
79. Khetselius YO (2015) Frontiers in quantum methods and applications in chemistry and
physics. In: Nascimento M, Maruani J, Brändas E, Delgado-Barrio E, Piecuch P (eds) Pro-
gress in theoretical chemistry and physics, vol 29. Springer, Cham, pp 55–76
80. Khetselius OYu (2008) Hyperfine structure of atomic spectra. Astroprint, Odessa
81. Khetselius OYu (2005) Hyperfine structure of radium. Photoelectronics. 14:83
82. Khetselius OYu (2012) J Phys Conf Ser 397:012012
83. Khetselius OYu, Florko TA, Svinarenko AA, Tkach TB (2013) Phys Scripta T153:014037
84. Khetselius OYu (2008) Spectral line shape. AIP Conf Proc 1058:363
85. Khetselius OYu, Glushkov AV, Gurnitskaya EP, Loboda AV, Mischenko EV, Florko TA,
Sukharev DE (2008) AIP Conf Proc 1058:231
86. Khetselius OYu (2007) Photoelectronics 16:129
87. Khetselius OYu, Gurnitskaya EP (2006) Sensor Electr Microsyst Techn N3, 35–39
88. Khetselius OYu, Gurnitskaya EP (2006) Sensor Electr and Microsyst Techn Issue 2:25−29
89. Feller D, Davidson ER (1981) J Chem Phys 74:3977
90. Froelich P, Davidson ER, Brändas E (1983) Phys Rev A 28:2641
91. Rittby M, Elander N, Brändas E (1983) Int J Quant Chem 23:865
92. Yan A, Wang C, Yung Y, Ya KC, Chen GH (2011) J Chem Phys 134:241103
93. Maruani J (2016) J Chin Chem Soc 63:33
94. Pavlov R, Mihailov L, Velchev Ch, Dimitrova-Ivanovich M, Stoyanov Zh, Chamel N,
Maruani J (2010) J Phys Conf Ser 253:012075
95. Dietz K, Heβ BA (1989) Phys Scripta 39:682
96. Kohn W, Sham LJ (1964) Phys Rev A 140:1133; Hohenberg P, Kohn W (1964) Phys Rev B
136:864
97. Gidopoulos N, Wilson S (2004) The fundamentals of electron density, density matrix and
density functional theory in atoms, molecules and the solid state. In: Progress in theoretical
chemistry and physics, vol 14. Springer, Berlin
98. Glushkov A V (2006) Relativistic and correlation effects in spectra of atomic systems.
Astroprint, Odessa
Relativistic Quantum Chemistry
and Spectroscopy of Exotic Atomic
Systems with Accounting for Strong
Interaction Effects
1 Introduction
It is well known that studying the energy, spectral, radiation parameters, including
the spectral lines hyperfine structure, for heavy exotic (hadronic, kaonic, pionic)
atomic systems is of a great interest for the further development as atomic and
nuclear theories and quantum chemistry of strongly interacted fermionic systems
(see, for example, Refs. [1–31]). Really, the exotic atoms enable to probe aspects of
atomic and nuclear structure that are quantitatively different from what can be
studied in the electronic (“usual”) atoms. Besides, the corresponding data on the
energy and spectral properties of the hadronic atomic systems can be used as a
powerful tool for the study of particles and fundamental properties.
At present time one of the most sensitive tests for the chiral symmetry breaking
scenario in the modern hadron’s physics is provided by studying the exotic
hadron-atomic systems. Nowadays the transition energies in pionic (kaonic, muonic
etc.) atoms are measured with an unprecedented precision and from studying
spectra of the hadronic atoms it is possible to investigate the strong interaction at
low energies measuring the energy and natural width of the ground level with a
precision of few meV [20–46].
The strong interaction is the reason for a shift in the energies of the low-lying
levels from the purely electromagnetic values and the finite lifetime of the state
corresponds to an increase in the observed level width. For a long time the similar
experimental investigations have been carried out in the laboratories of Berkley,
Virginia (USA), CERN (Switzerland).
The most known theoretical models to treating the hadronic (pionic, kaonic,
muonic, antiprotonic etc.) atomic systems are presented in Refs. [1–46]. The most
difficult aspects of the theoretical modelling are reduced to the correct description of
pion-nuclear strong interaction [10–18] as the electromagnetic part of the problem
can be in principle reasonably accounted for [47–60].
In the present chapter we briefly present the fundamentals of a consistent rela-
tivistic theory of spectra of the exotic pionic atomic systems (with simultaneous
accounting for the electromagnetic and strong pion-nuclear interactions by means of
using the generalized radiation and strong pion-nuclear optical potentials) on the
basis of the Klein-Gordon-Fock. The nuclear and radiative corrections are effec-
tively taken into account. The modified Uehling-Serber approximation is used to
take into account for the Lamb shift polarization part. In order to take into account
the contribution of the Lamb shift self-energy part we have used the generalized
non-perturbative procedure, which generalizes the Mohr procedure and radiation
model potential method by Flambaum-Ginges. There are presented data of calcu-
lation of the energy and spectral parameters for pionic atoms of the 93Nb, 173Yb,
181
Ta, 197Au, with accounting for the radiation (vacuum polarization), nuclear
(finite size of a nucleus) and the strong pion-nuclear interaction corrections. The
measured values of the Berkley, CERN and Virginia laboratories and alternative
data based on other versions of the Klein-Gordon-Fock theories with taking into
Relativistic Quantum Chemistry and Spectroscopy of Exotic … 73
account for a finite size of the nucleus in the model uniformly charged sphere and
the standard Uhling-Serber radiation corrections are listed too.
Here we present a brief description of the key moments of our approach (more
details can be found in Refs. [61–79]). The relativistic electron wave functions are
determined from solution of the Klein-Gordon-Fock equation (pion is the Boson
with spin 0, mass: mπ − = 139.57018 MэB, rπ− = 0.672 ± 0.08 fm) with a general
potential (the latter includes an electric and polarization potentials of a nucleus plus
the strong pion-nuclear interaction potential), which can be written as follows:
1
m2 c2 ΨðxÞ = f ½iℏ∂t + eV0 ðrÞ2 + ℏ2 ∇2 gΨðxÞ ð1Þ
c2
where c is a speed of the light, h is the Planck constant, and Ψ0(x) is the scalar wave
function of the space-temporal coordinates. Usually one considers the central
potential [V0(r), 0] approximation with the stationary solution:
1
f ½E + eV0 ðrÞ2 + ℏ2 ∇2 − m2 c2 gφðxÞ = 0 ð3Þ
c2
Here E is the total energy of the system (sum of the mass energy mc2 and
binding energy ε0).
In principle, the central potential V0 is the sum of the following potentials: the
electric potential of a nucleus, vacuum-polarization potential and the strong inter-
action potential. The nuclear potential for the spherically symmetric density ρðrjRÞ
can be presented as follows:
Zr 0 Z∞ 0 0 0
dr r ρ r R + dr r ρ r R
0 0
Vnucl ðrjRÞ = − ð1 ̸ r Þ 2
ð4Þ
0 r
̸2
pffiffiffi
ρðrjRÞ = 4γ3 ̸ π exp − γr 2 ð5Þ
Z∞
drr 2 ρðrjRÞ = 1,
0
Z∞
drr 3 ρðrjRÞ = R,
0
(here g = 4/R2 and R is the effective nucleus radius) or by the Fermi function:
where the parameter a = 0.523 fm, the parameter c is chosen by such a way that it
is true the following condition for average-squared radius:
Further one should use the formulas for the finite size nuclear potential and its
derivatives on the nuclear radius. Here we use the known Ivanov-Ivanova et al.
method of differential equations (look details in Refs. [80–83]). The effective
algorithm for definition of the potential Vnucl ðrjRÞ is used in Refs. [65, 72] and
reduced to solution of the following system of the differential equations (for the
Fermi model):
Zr 0
0 0 0
V nuclðr, RÞ = 1 ̸ r 2 dr r 2 ρ r , R ≡ 1 ̸ r 2 yðr, RÞ,
0
with the corresponding boundary conditions. In a case of the Gaussian model the
corresponding system of differential equations is as follows:
Rr
V ′ nuclðr, RÞ = ð1 ̸ r 2 Þ dr ′ r ′2 ρ r ′ , R ≡ ð1 ̸r 2 Þyðr, RÞ
0 ð9Þ
y′ ðr, RÞ = r 2 ρðr, RÞ
pffiffiffi 8r
ρ′ ðr, RÞ = − 8γ5 ̸2 r ̸ π exp − γr 2 = − 2γrρðr, RÞ = − 2 ρðr, RÞ
πr
yð0, RÞ = 0,
̸2
pffiffiffi
ρð0, RÞ = 4γ3 ̸ π = 32 ̸ R3 ð10Þ
(A4)
(A5)
76 O. Yu. Khetselius et al.
QED effect of the vacuum polarization: A1—the Uehling-Serber term; A2, A3–
terms of order члeны пopядкa [αðZαÞn (n = 2, …); A4- the Källen-Sabry cor-
rection of order α2 ðαZ Þ; A5–the Wichmann-Kroll correction of order кa αðZαÞn
(n = 3). An effect of the vacuum polarization is usually taken into account in the
first PT theory order by means of the generalized Uehling-Serber potential with
modification to take into account the high-order radiative corrections. In particular,
the generalized Uehling-Serber potential can be written as follows:
Z∞ pffiffiffiffiffiffiffiffiffiffiffi
2α t2 − 1 2α
U ðr Þ = − dt expð − 2rt ̸ αZ Þ 1 + 1 ̸2t 2
≡− C ðgÞ, ð11Þ
3πr t 2 πr
1
where g = r ̸ ðαZÞ. More correct and consistent approach is presented in Refs. [42,
43, 52–62]. An accounting of the nuclear finite size effect modifies the potential (7)
as follows:
Z Z∞ pffiffiffiffiffiffiffiffiffiffiffi
2α2 1 t2 − 1 ρ r′
U ðr Þ = −
FS 3 ′
d r dt exp − 2t r − r ′ ̸αZ × 1+ 2 ,
3π 2t t 2 jr − r ′ j
m
ð12Þ
Z4
ESE ðHjZ, nljÞ = 0.027148 F ðHjZ, nljÞ ð13Þ
n3
The values of F are given at Z = 10 − 110, nlj = 1s, 2s, 2p1 ̸2 , 2p3 ̸2 .
These results are modified here for the states 1 s2 nlj of the non-H atoms (ions).
It is supposed that for any ion with nlj electron over the core of closed shells the
sought value may be presented in the form [52]:
Relativistic Quantum Chemistry and Spectroscopy of Exotic … 77
ξ4
ESE ðZ, nljÞ = 0.027148 3
f ðξ, nljÞ cm − 1 ð14Þ
n
The most difficult aspect of the problem is an adequate account for the strong
pion-nuclear interaction in the exotic system. Now it is well known that the most
fundamental and consistent microscopic theory of the strong interactions is pro-
vided by the modern quantum chromodynamics. One should remind that here
speech is about a gauge theory based on the representation of the confined coloured
quarks and gluons. Naturally one could consider the regimes of relatively low and
high energies (asymptotic freedom). In a case of the low energies so called coupling
constant increases to the order 1 and, therefore, this perturbation methods fail to
describe the interaction of strongly interacting hadrons (including pions). Naturally,
to describe the strong pion-nuclear interaction (even at relatively low energies)
microscopically, a different approaches can be developed (look details in Refs. [11–
19, 76–79]).
More simplified and sufficiently popular approach to treating the strong inter-
action in the pionic atomic system is provided by the well known optical potential
model (c.g. [14, 15]). On order to describe the strong π−N interaction we have used
the optical potential model n which the generalized Ericson-Ericson potential is as
follows:
78 O. Yu. Khetselius et al.
4π αðr Þ
Vπ − N = Vopt ðr Þ = − qðr Þ∇ ∇ ,
2m 1 + 4 ̸3πξαðr Þ
mπ
qð r Þ = 1 + b0 ρðr Þ + b1 ρn ðr Þ − ρp ðr Þ
mN
mπ
+ 1+ B0 ρ2 ðr Þ + B1 ρðr Þδρðr Þ , ð15Þ
2mN
mπ − 1
αðr Þ = 1 + c 0 ρð r Þ + c 1 ρn ð r Þ − ρp ð r Þ
mN
mπ − 1
+ 1+ C0 ρ2 ðr Þ + C1 ρðr Þδρðr Þ .
2mN
Further we note that an energy of the hadronic atom can be represented as the
following sum:
Here EKG -is the energy of a pion in a nucleus ðZ, AÞ with the point-like charge
(dominative contribution in (17)), EFS is the contribution due to the nucleus finite
size effect, EVP is the radiation correction due to the vacuum-polarization effect, EN
is the energy shift due to the strong interaction VN .
It is easily to note that the strong pion-nucleus interaction contribution into
energy can be directly found from the solution of the Klein-Gordon-Fock equa-
tion with the corresponding pion-nucleon potential, for example, in the optical
potential approximation (8). Since the corresponding optical potential contains the
complex parameters, the relevant energy eigen-values of the Klein-Gordon-Fock
Relativistic Quantum Chemistry and Spectroscopy of Exotic … 79
equation for the definite pionic state ði = nlÞ in an atom are the complex values
too, i.e. [11, 38, 76]:
where the imaginary part determines a width of pionic energy level Gi. The total
width of any level is determined as by the strong pion-nuclear interaction contri-
bution ΓSi (pion absorption) as by the electromagnetic contribution Γrad
i . The latter is
determined by a probability of the electromagnetic radiation transition (including
the Auger process probability ΓAi ) on the lower level. As an example, for the width
of pion 1 s can be written:
ΓS1s = Γexp rad S A exp
2p → 1s − Γ2p + Γ2p + Γ2p ≈ Γ2p → 1s , ð19Þ
where Γradi i ΓAi are the radiation and Auger widths respectively. Let us consider
further elements of theory, associated with the implementation of the known rel-
ativistic energy formalism in our theory to calculate the electromagnetic interaction
transition probabilities in spectrum of the pionic atom [80–91]. It is worth to remind
that in relativistic theory of the usual many-electron systems (an energy of any
excited state is a complex quantity) an shift of the total energy level is usually
represented as:
where Γradi is a radiation width, and the corresponding radiative transition proba-
bility in the usual atomic system P ∼ Γrad
i . In order to compute the latter we use the
generalized relativistic energy approach.
Let us remind that an initial general energy formalism combined with an
empirical model potential method in a theory of atoms and multicharged ions has
been developed by Ivanov-Ivanova et al. [80–84]; further more general ab initio
gauge-invariant version of relativistic energy approach has been presented by
Glushkov-Ivanov [89]. The imaginary part of the energy shift of an atom is con-
nected with the radiation decay possibility (transition probability). For the α-n
radiation transition ImDE in the lowest order of the PT is determined as:
1
Im ΔE = − ∑ V jωαn j , ð21Þ
4π α > n > f αnαn
½α < n ≤ f
where ωαn is a frequency of the α-n radiation, (α > n > f) for particle and (α < n <
f) for vacancy. The matrix element V is determined as follows:
80 O. Yu. Khetselius et al.
ZZ
jωj sinjωjr12
Vijkl = dr1 dr2 ψ *i ðr1 Þψ *j ðr2 Þ ð1 − α1 α2 Þψ *k ðr2 Þψ *l ðr1 Þ ð22Þ
r12
The detailed procedure for computing the matrix elements (22) is presented in
Refs. [76–92]. All calculations are performed with using the numeral code
Superatom (version 98).
Table 2 The values of some optical potential parameters, which have been used in different
calculations (see text)
Tauscher Tauscher Batty Seki Nagels Row78 Laat Odes
ξ=0 ξ=1 ξ=1 ξ=1 ξ=1 ξ=1 ξ=1 ξ = 1 [76,
[16] [16] [17] [18] [22] [19] [38, 39] 79]
b0 −0.0296 −0.0293 −0.017 0.003 −0.013 −0.004 0.007 0.003 mπ− 1
b1 −0.077 −0.078 −0.13 −0.143 −0.092 −0.094 −0.075 −0.094 mπ− 1
ReB0 0 0 −0.0475 −0.15 – – −0.18 −0.15 mπ− 4
ImB0 0.0436 0.0428 0.0475 0.046 – – 0.058 0.046 mπ− 4
c0 0.172 0.227 0.255 0.21 0.209 0.23 0.266 0.21 mπ− 3
c1 0.22 0.18 0.17 0.18 0.177 0.17 0.40 0.18 mπ− 3
ReC0 – – – 0.11 – – 0.07 0.11 mπ− 6
ImC1 – – – – – – −0.34 −0.25 mπ− 6
as Ta, Bi et al.) have shown that the appropriate values of width due to strong
interaction by a factor two and more are less than the values specified within the
optical potential model with using the earliest parameterization [14, 15].
Such relatively significant deviations are typical for strong levels shifts.
A similar anomaly was detected in a case of the strong quadrupole shift of the
pion energy in 3d and 4f states. It should be noted that the results of the hyperfine
structure studying for heavy pionic atoms with pion at low (deep) lying states are
extremely scarce.
Earlier it is indicated a possibility of improving the consistency between theory
and experiment by taking into account of the strong interaction potential isovector
(absorption) terms. In fact, it is a more adequate account of increasing S-wave
repulsion for absorption of pion at 3d-level. In this regard, under the parameteri-
zation of the optical potential authors [33] left without changing the parameters
settings that are the most reliably identified, namely: ReB0, ImB0, c0, c1, ReC0,
ImC0.
At the same time the parameters whose values differ most strongly in different
sets, in particular, b1 (plus parameters, which are usually not included so far in the
basic optical potential parameterizations, i.e. ImB1, ImC1), should be optimized.
This is achieved by receiving preliminary calculated relationships (illustrations are
given below) for the energy shifts and widths (for a number of states fof the
following systems: 20Ne, 24Mg, 93Nb, 133Cs, 175Lu, 181Ta, 197Au, 208Pb) upon the
b1, ImB1, ImC1 parameter values; further, there were chosen such values that satisfy
the smallest standard deviation of the experimental values.
In Tables 3 and 4 we list the dependences of shifts and widths for the 4f, 3d
levels due to the strong pion nuclear interaction upon the parameter ImB1 ImC1
values for the pionic atom of 208Pb (our data).
Further in Tables 5 and 6 we list the analogous dependences of shifts and widths
for the 4f, 3d levels due to the strong pion nuclear interaction upon the parameter
ImB1 ImC1 values for the pionic system 181Ta.
82 O. Yu. Khetselius et al.
Table 3 The dependence of shifts and widths for the 4f, 3d levels due to the strong pion nuclear
interaction upon the parameter ImB1 value for 208Pb
ImB1 ε4f
0 (
208
Pb) Γ4f
0 (
208
Pb) ε3d
0 (
208
Pb) Γ3d
0 (
208
Pb)
0.00 1.596 1.11 18.2 67.7
0.02 1.603 1.09 18.9 63.6
0.04 1.625 1.07 19.7 58.7
0.06 1.633 1.05 20.5 54.8
0.08 1.641 1.03 21.6 50.5
0.10 1.652 0.96 22.8 47.2
0.12 1.723 0.91 23.8 45.3
Exp. 1.68 ± 0.04 0.98 ± 0.05 22.7 ± 2.2 47.1 ± 3.6
Table 4 The dependence of shifts and widths for the 4f, 3d levels due to the strong pion nuclear
interaction upon the parameter ImC1 для ядpa 208Pb
ImC1 ε4f
0 (
208
Pb) Γ4f
0 (
208
Pb) ε3d
0 (
208
Pb) Γ3d
0 (
208
Pb)
0.00 1.610 1.050 18.8 68.3
−0.05 1.622 0.949 19.4 64.4
−0.10 1.633 0.942 19.8 60.3
−0.15 1.642 0.939 20.5 56.5
−0.20 1.653 0.928 21.8 52.3
−0.25 1.665 0.918 22.6 48.2
−0.30 1.676 0.899 23.2 44.5
Exp. 1.68 ± 0.04 0.98 ± 0.05 22.7 ± 2.2 47.1 ± 3.6
Table 5 The dependence of shifts and widths for the 4f, 3d levels due to the strong pion nuclear
interaction upon the parameter ImB1 value for 208Pb 181Ta
ImB1 ε4f
0 (
181
Ta) Γ4f
0 (
181
Ta) ε3d
0 (
181
Ta) Γ3d
0 (
181
Ta)
0.00 0.508 0.41 13.5 34.9
0.02 0.517 0.39 14.1 30.8
0.04 0.525 0.37 14.7 27.5
0.06 0.533 0.35 15.2 24.8
0.08 0.543 0.33 15.8 22.6
0.10 0.554 0.30 16.3 20.3
0.12 0.566 0.28 16.8 18.3
Exp. 0.56 ± 0.04 0.31 ± 0.05 16.2 ± 1.3 20.1 ± 1.5
3.2 Pionic Hydrogen and the Radiation Widths of the 3d, 4f,
5g Levels in Some π − A Systems
In Table 7 we list the calculated (in meV) QED corrections to the energies of the 1s,
2p, 3p, 4p states for pionic hydrogen: data by Indelicato et al. and Serga et al. and
Relativistic Quantum Chemistry and Spectroscopy of Exotic … 83
Table 6 The dependence of shifts and widths for the 4f, 3d levels due to the strong pion nuclear
interaction upon the parameter ImC1 для ядpa 81Ta
ImC1 ε4f 81
0 ( Ta) Γ4f 81
0 ( Ta)
ε3d 81
0 ( Ta) Γ3d 81
0 ( Ta)
0.00 0.486 0.334 12.8 33.7
−0.05 0.502 0.325 13.7 30.8
−0.10 0.519 0.313 14.3 28.5
−0.15 0.534 0.302 14.9 25.7
−0.20 0.547 0.274 15.6 23.4
−0.25 0.559 0.268 16.3 20.6
−0.30 0.571 0.255 16.8 19.9
Exp. 0.56 ± 0.04 0.31 ± 0.05 16.2 ± 1.3 20.1 ± 1.5
our data. The following abbreviations are used: the Uehling-Serber vacuum
polarization correction (VP-US), the Kallen-Sabry correction (VP-KS),
Wichman-Kroll one (PV-WK).
Analysis of data shows that there is physically reasonable agreement between
different theoretical data, namely, data by Schlesser-Indelicato et al. [31, 32, 76–
79]. Naturally, the reason is obvious as the known expansion parameter aZ in the
hydrogen atom is significantly less than one, and the general QED contributions
into the levels energies are not large in comparison with other ones.
In Table 8 we present our calculated (the relativistic Klein-Gordon-Fock theory
combined with energy approach) data on the radiation widths of the 3d, 4f, 5g
levels for a number of the pionic π−A atoms. There are also listed the analogous
Table 7 The calculated (in QED contr. 1s [31, 32] 1s [76, 77] 1s [79]
meV) QED corrections to the
energies of the 1s, 2p, 3p, 4p VP-US −3240.802 −3240.799 −3240.801
states for pionic hydrogen: VP-KS −24.365 −24.363 −24.365
data by Indelicato et al., Serga PV-WK −4.110 −4.113 −4.112
et al. and our data QED contr. 2p 2p 2p
VP-US −35.795 −35.793 −35.794
VP-KS −0.346 −0.343 −0.345
PV-WK −0.008 −0.010 −0.009
QED contr. 3p 3p 3p
VP-US −11.407 −11.405 −11.406
VP-KS −0.108 −0.105 −0.107
PV-WK −0.002 −0.003 −0.002
QED contr. 4p 4p 4p
VP-US −4.921 −4.918 −4.920
VP-KS −0.046 −0.044 −0.045
PV-WK −0.001 −0.002 −0.001
84 O. Yu. Khetselius et al.
Table 8 The radiation widths of the 3d, 4f, 5g levels in some π − A systems
Nucleus Γ1rad Γ2rad (5g) Γ1rad (4f) Γ2rad (4f) [79] Γ1rad (3d) Γ2rad (3d)
(5g) [38, 39] [79] [38, 39] [38, 39] [79]
165
Ho – 15.2 – 56.1 – 228.8
173
Yb – 17.9 – 66.8 – 275.4
175
Lu – 20.7 – 77.5 – 320.3
181
Ta 25.7 23.5 90.9 88.6 369.9 366.1
203
Tl – 37.2 – 136.8 – 557.2
208
Pb 41.5 39.4 146.8 143.2 587.6 583.8
209
Bi 43.7 41.5 156.2 153.1 617.3 613.7
data by Laat-Konijn et al. [38, 39], obtained with using the relativistic
Klein-Gordon-Fock model and Hartree-Fock approximation.
In Figs. 2 and 3 there are presented the parts of the X-ray spectra of the 165Ho
(Fig. 2), 181Ta (Fig. 3) and positions of the hyperfine structure components (5 g-4f
transition; experimental data from [13]).
In Tables 9a and 10 we present theoretical and experimental data (in keV) for the
4f level shift (a) and widths (b) provided by the strong pion-nuclear interaction for a
number of pionic atoms. The shortened designation of the parameter sets for the
strong π−N interaction potential: Tauscher—Tau1; Tauscher, Tau2; Batty et al.—
Bat; Seki et al.—Sek; de Laat-Konijin et al.—Laat, this work—Srg-Sha [16–22, 38,
39, 76–79]. In our parameterization of the strong p−N interaction potential the most
reliably defined (B0, c0, c1, C0) parameters are remained unchanged, and the
parameters whose values differ greatly in different sets, in particular, b1
(b1 = −0.094) plus still not included ones ImB1, ImC1 have been optimized by
computing dependencies of the strong shifts upon the parameters b1, ImB1, ImC1
for π−−20Ne, 24Mg, 93Nb, 133Cs,175Lu, 181Ta, 197Au, 208Pb atoms. Further we have
chosen the values which satisfy the smallest standard deviation of reliable experi-
mental values.
In Table 11 the analogous theoretical and experimental data (in keV) for the 3d
level shift (a) and widths for different pionic atoms are listed [16–22, 38, 39,
76–79].
In Table 12 data on the 4f-3d, 5g-4f transition energies for pionic atoms of the
93
Nb, 173Yb, 181Ta, 197Au are presented. There are also listed the measured values
Relativistic Quantum Chemistry and Spectroscopy of Exotic … 85
of the Berkley, CERN and Virginia laboratories and alternative data obtained on the
basis of computing within alternative versions of the Klein-Gordon-Fock
(KGF) theory with taking into account for a finite size of the nucleus in the
model uniformly charged sphere and the standard Uehling-Serber radiation cor-
rection (see Refs. [6, 7, 13, 42, 43, 86, 89]).
The analysis of the presented data indicate on the importance of the correct
accounting for the radiation (vacuum polarization) and the strong pion-nuclear
interaction corrections. The contributions due to the nuclear finite size effect should
be accounted in a precise theory too. More exact knowledge of the electromagnetic
interaction parameters for a pionic atom will make more clear the true values for
parameters of the pion-nuclear potentials. Further it allows to correct a disadvantage
of widely used parameterization of the optical potential. It is especially important if
one takes into account an increasing accuracy of the X-ray pionic atom spec-
troscopy experiments. It is interesting to note that the contributions into transition
energies are about ∼5 keV due to the QED effects, ∼0.2 keV due to the nuclear
86 O. Yu. Khetselius et al.
Table 9 Theoretical and experimental data for the 4f level shift (keV) provided by the strong
pion-nuclear interaction for a number of pionic atoms (see text)
ε4f , Γ4f Exp H-like Tau 1 Tau 2 Ba ξ = 1 Sek ξ = 1 Laat ξ = 1 Srg-Sha Ou r
Func. ξ=0 ξ=1 ξ=1 ξ=1
165
Ho: ε 0.29 ± 0.01 0.21 0.25 0.24 0.24 0.21 0.26 0.29 0.29
0.27 0.26
169
Tm: ε – – – – – – – 0.38 0.38
173
Yb: ε – – – – – – – 0.44 0.44
175
Lu: ε 0.51 ± 0.04 0.36 0.43 0.42 0.41 0.36 0.46 0.50 0.50
181
Ta: ε 0.56 ± 0.04 0.47 0.57 0.54 0.53 0.47 0.60 0.55 0.55
197
Au: ε 1.25 ± 0.07 – 1.21 1.14 1.12 0.98 1.25 1.24 1.24
208
Pb: ε 1.68 ± 0.04 – 1.76 1.62 1.58 1.39 1.68 1.65 1.65
209
Bi: ε 1.78 ± 0.06 – 1.94 1.80 1.78 1.57 1.83 1.77 1.77
finite size effect, and ∼0.07 keV due to the electron screening effect, provided by
the 2[He], 4[Be], 10[Ne] electron shells [79].
To conclude, let us underline that the key factors for the physically reasonable
agreement between experimental and theoretical data on the multi-electron pionic
Relativistic Quantum Chemistry and Spectroscopy of Exotic … 87
Table 10 Theoretical and experimental data for the 4f level widths (keV) provided by the strong
pion-nuclear interaction for a number of pionic atoms (see text)
ε4f , Γ4f Exp H-like Tau 1 Tau 2 Bat Sek Laat Srg-Sha Our
Func. ξ=0 ξ=1 ξ=1 ξ=1 ξ=1 ξ=1 ξ=1
165
Ho: Γ 0.21 ± 0.02 0.08 0.13 0.12 0.13 0.11 0.13 0.20 0.21
169
Tm: Γ – – – – – – – – 0.23
173
Yb: Γ – – – – – – – – 0.26
175
Lu: Γ 0.27 ± 0.07 0.14 0.23 0.22 0.24 0.20 0.24 0.28 0.28
181
Ta: Γ 0.31 ± 0.05 0.16 0.31 0.30 0.31 0.27 0.31 0.30 0.31
197
Au: Γ 0.77 ± 0.04 – 0.73 0.68 0.69 0.58 0.67 0.75 0.77
208
Pb: Γ 0.98 ± 0.05 – 1.18 1.04 1.03 0.86 0.98 0.97 0.99
209
Bi: Γ 1.24 ± 0.09 – 1.35 1.18 1.17 0.99 1.10 1.22 1.25
Table 11 Theoretical and experimental data for the 3d level shifts and widths (keV) provided by
the strong pion-nuclear interaction for a number of pionic atoms (see text)
3d Еxp. Tau 1 Tau 2 Bat Sek Laat Srg-Sha Odes
ξ=0 ξ=1 ξ=1 ξ=1 ξ=1 ξ=1 ξ=1
93
Nb: ε 0.74 ± 0.02 0.66 0.67 0.73 0.66 0.75 0.75 0.73
169
Tm: ε – – – – – – – 11.0
173
Yb: ε – – – – – – – 12.4
175
Lu: ε – – – – – – – 13.9
181
Ta: ε 16.2 ± 1.3 19.6 16.4 10.4 4.4 14.4 16.3 16.1
197
Au: ε 20.6 ± 1.9 27.9 22.5 13.2 5.0 20.3 21.1 20.3
208
Pb: ε 22.7 ± 2.2 34 25 13 3 18 22.8 22.6
209
Bi: ε 20 ± 3 37 27 17 5 19 23 21.1
93
Nb: Γ 0.40 ± 0.02 0.405 0.413 0.459 0.404 0.452 0.42 0.41
169
Tm: Γ – – – – – – – 15.7
173
Yb: Γ – – – – – – – 17.6
175
Lu: Γ – – – – – – – 19.4
181
Ta: Γ 20,1 ± 1,5 40,5 37,5 33,4 26,2 27,6 20.3 20.2
197
Au: Γ 34±3.6 68 62 53 41 42 36.2 35.6
208
Pb: Γ 47.1 ± 3.6 88 78 65 51 51 47.2 47.0
209
Bi: Γ 52 ± 3.6 97 86 72 57 56 53.6 53.4
Table 12 Transition energies (keV) in the spectra of some heavy pionic atoms (see text)
π−A Trans. Berkley CERN EKGF+EM EKGF-EM EN [6] EN EN [89]
Eexp Eexp [6, 7] [13] [13] a
93
Nb 5g-4f – 307.79 ± – – – – 307.85
0.02
173
Yb 5g-4f – – – – – – 412.26
181
Ta 5g-4f 453.1 ± 453.90 ± 453.06 453.78 – 453.52 453.71
0.4 0.20 453.62
197
Au 5g-4f 532.5 ± 533.16 ± 528.95 – 532.87 531.88 533.08
0.5 0.20
93
Nb 4f-3d – 140.3 ± – – – – 140.81
0.1
173
Yb 4f-3d – – – – – – 838.67
181
Ta 4f-3d – 1008.4 ± – – – 992.75 1008.80
1.3
References
16. Tauscher L (1971) Analysis of pionic atoms and the p-nucleus optical potential. In:
Proceedings of the international semantic p-Meson nucleus interaction-CNRS-strasbourg,
France, p 45
17. Batty C, Biagi S, Friedman E, Hoath S (1983) Phys Rev Lett 440:931; Batty C J, Friedman E,
Gal A (1978) Nucl Phys A 402:411
18. Seki R, Masutani K, Jazaki K (1983) Phys Rev C 27:1817
19. Rowe G, Salamon M, Landau RH (1978) Phys Rev C 18:584
20. Anagnostopoulos D, Biri S, Boisbourdain V, Demeter M, Borchert G et al. (2003) Nucl Inst
Meth B 205:9
21. Anagnostopoulos D, Gotta D, Indelicato P, Simons LM (2003) arXiv:physics. 0312090v1
22. Nagels MM, de Swart J, Nielsen H et al (1976) Nucl Phys B 109:1
23. Lauss B (2009) Nucl Phys A 827C, 401 PSI experiment R-98.01http://pihydrogen.psi.ch
24. CERN DIRAC Collaboration (2011) Search for long-lived states of p+ p− and p−K atoms,
CERN-SPSLC-2011–001 SPSLC-P-284-ADD p 22
25. Umemoto Y, Hirenzaki S, Kume K, Toki H, Tahihata I (2001) Nucl Phys A 679:549
26. Nose-Togawa N, Hirenzaki S, Kume K (1999) Nucl Phys A 646:467
27. Glushkov AV, Malinovskaya SV, Gurnitskaya EP, Khetselius OYu, Dubrovskaya YuV
(2006) J Phys: Conf Ser 35:425
28. Hatsuda T, Kunihiro T (1994) Phys Rep 247:221
29. Ikeno N, Kimura R, Yamagata-Sekihara J, Nagahiro H, Jido D, Itahashi K et al.
(2011) 1107.5918v1[nucl-th]
30. Kolomeitsev EE, Kaiser N, Weise W (2003) Phys Rev Lett 90:092501
31. Lyubovitskij V, Rusetsky A (2000) Phys Lett B 494:9
32. Schlesser S, Le Bigot E-Q, Indelicato P, Pachucki K (2011) Phys Rev C 84:015211
33. Sigg D, Badertscher A, Bogdan M, Goudsmit P, Leisi H, Schröder H, Zhao Z, Chatellard D,
Egger J, Jeannet E, Aschenauer E, Gabathuler K, Simons L (1996) Rusi El Hassan A. Nucl
Phys A 609:269
34. Gotta D, Amaro F, Anagnostopoulos D, Biri S, Covita D, Gorke H, Gruber A, Hennebach M,
Hirtl A, Ishiwatari T, Indelicato P, Jensen T, Bigot E, Marton J, Nekipelov M, dos Santos J,
Schlesser S, Schmid P, Simons L, Strauch H, Trassinelli M, Veloso J, Zmeskal J ed. Kania Y,
Yamazaki Y (AIP) (2008) CP1037, 162
35. Gotta D, Amaro F, Anagnostopoulos D, Biri S, Covita D, Gorke H, Gruber A, Hennebach M,
Hirtl A, Ishiwatari T, Indelicato P, Jensen T, Bigot E, Marton J, Nekipelov M, dos Santos J,
Schlesser S, Schmid P, Simons L, Strauch H, Trassinelli M, Veloso J, Zmeskal (2008)
Precision physics of simple atoms and molecules. In: Lecture notes in physics, vol 745,
Springer, Berlin, Heidelberg, pp 165–186
36. Taal A, D’Achard van Enschut J, Berkhput J et al (1985) Phys Lett B 156:296
37. Taal A, David P, Hanscheid H, Koch JH, de Laat CT et al (1990) Nucl Phys A 511:573
38. de Laat CT, Taal A, Konijn J et al (1991) Nucl Phys A 523:453
39. de Laat CT, Taal A, Duinker W et al (1987) Phys Lett B 189:7
40. Khetselius OYu, Turin AV, Sukharev DE, Florko TA (2009) Sensor Electr and Microsyst
Techn N1, 30–35
41. Mohr PJ (1993) Atom Dat Nucl Dat Tabl 24:453; (1983) Phys Scripta 46:44
42. Indelicato P (1996) Phys Scripta T65:57; Indelicato P, Trassinelli M (2005) arXiv:
physics.0510126v1
43. Glushkov AV, Malinovskaya SV (2003) In: Fazio G, Hanappe F (eds) New projects and new
lines of research in nuclear physics. World Sci, Singapore, pp 242–250
44. Santos J, Parente F, Boucard S, Indelicato P, Desclaux J (2005) Phys Rev A 71:032501
45. Mitroy J, Ivallov IA (2001) J Phys G Nucl Part Phys 27:1421
46. Strauch T (2009) High-precision measurement of strong-interaction effects in pionic
deuterium, Julich
47. Serot B, Walecka J (1986) Relativistic nuclear many body problem. In: Advances in nuclear
physics, Plenum Press, N-Y
48. Glushkov AV, Ivanov LN (1992) Phys Lett A 170:33
90 O. Yu. Khetselius et al.
49. Glushkov AV, Ivanov LN, Ivanova EP (1986) Autoionization phenomena in atoms. Moscow
University Press, Moscow, pp 58–160
50. Ivanova EP, Glushkov AV (1986) J Quant Spectr Rad Transfer 36:127
51. Glushkov AV, Khetselius OYu, Gurnitskaya EP, Loboda AV, Florko TA, Sukharev DE,
Lovett L (2008) Frontiers in quantum systems in chemistry and physics. In: Wilson S,
Grout PJ, Maruani J, Delgado-Barrio G, Piecuch P (eds) Progress in Theoretical Chemistry
and Physics, vol 1. Springer, Dordrecht, pp 507–524
52. Flambaum VV, Ginges JSM (2005) Phys Rev A 72:052115
53. Safranova UI, Safranova MS, Johnson WR (2005) Phys Rev A 71:052506
54. Glushkov AV, Malinovskaya SV, Khetselius OYu, Loboda AV, Sukharev DE, Lovett L
(2009) Int J Quant Chem 109:1717
55. Gurnitskaya EP, Khetselius OYu, Loboda AV, Vitavetskaya LA (2008) Photoelectronics
17:127
56. Glushkov AV, Ambrosov SV, Loboda AV, Gurnitskaya EP, Prepelitsa GP (2005) Int J Quant
Chem 104:562
57. Johnson W, Sapirstein J, Blundell S (1993) Phys Scripta T 46:184
58. Glushkov AV (2006) Relativistic and correlation effects in spectra of atomic systems,
Astroprint, Odessa
59. Glushkov AV, Khetselius OYu, Malinovskaya SV (2008) Europ Phys Journ ST 160:195
60. Buyadzhi VV, Glushkov AV, Lovett L (2014) Photoelectronics 23:38
61. Glushkov AV, Ambrosov SV, Loboda AV, Chernyakova GYu, Svinarenko AA, Khet-
selius OY (2004) Nucl Phys A Nucl Hadr Phys 734:21
62. Glushkov AV, Ambrosov SV, Loboda AV, Gurnitskaya EP, Khetselius OY (2006) Recent
advances in the theory of chemical and physical systems. In: Julien P, Maruani J, Mayou D,
Wilson S, Delgado-Barrio G (eds) Progress in theoretical chemistry and physics, vol 15.
Springer, Dordrecht, pp 285–299
63. Glushkov AV (2005) AIP Conf Proceedings 796 (1): 206–210
64. Glushkov AV, Lovett L, Khetselius OYu, Gurnitskaya EP, Dubrovskaya YuV, Loboda AV
(2009) Int J Mod Phys A 24:611
65. Glushkov AV, Khetselius OY and Lovett L (2009) Advances in the theory of atomic and
molecular systems. Dynamics, spectroscopy, clusters and nanostructures. In: Piecuch P,
Maruani J, Delgado-Barrio G, Wilson S (eds) Progress in theoretical chemistry and physics,
vol 20. Springer, Dordrecht, pp 125–152
66. Khetselius OY (2012) Quantum systems in chemistry and physics. In: Nishikawa K,
Maruani J, Brändas E, Delgado-Barrio G, Piecuch P (eds) Progress in theoretical chemistry
and physics, vol 26. Springer, Dordrecht, pp 217–229
67. Khetselius OYu (2009) Int J Quant Chem 109:3330
68. Khetselius OYu (2009) Phys Scripta T135:014023
69. Khetselius OYu (2015) Frontiers in quantum methods and applications in chemistry and
physics. In: Nascimento M, Maruani J, Brändas E, Delgado-Barrio G (eds) Progress in
theoretical chemistry and physics, vol 29. Springer, Cham, pp 55–76
70. Khetselius OYu (2008) Hyperfine structure of atomic spectra. Astroprint, Odessa
71. Glushkov AV, Khetselius OYu, Svinarenko AA (2013) Phys Scripta T153:014029
72. Glushkov AV, Khetselius OY, Svinarenko AA (2012) Advances in the theory of quantum
systems in chemistry and physics. In: Hoggan P, Brändas E, Maruani J, Delgado-Barrio G,
Piecuch P (eds) Progress in theoretical chemistry and physics, vol 22. Springer, Dordrecht, pp
51–68
73. Khetselius O, Florko T, Svinarenko A, Tkach T (2013) Phys Scripta T153: 014037
74. Khetselius OYu (2010) AIP Conf Proc 1290:29
75. Khetselius OYu, Florko TA, Nikola LV, Svinarenko AA, Serga IN, Tkach TB,
Mischenko EV (2010) Quantum Theory: reconsideration of foundations (AIP). 1232:243
76. Serga IN, Dubrovskaya YV, Kvasikova AS, Shakhman AN, Sukharev DE (2012) J Phys Conf
Ser 397:012013
77. Serga IN (2013) Photoelectronics 22:71; Shakhman AN (2015) Photoelectronics 24:109
Relativistic Quantum Chemistry and Spectroscopy of Exotic … 91
78. Sukharev DE, Khetselius OYu and Dubrovskaya YuV (2009) Sensor Electr and Microsyst
Techn N3, 16–21
79. Bystryantseva AN, Khetselius OYu, Dubrovskaya YuV, Vitavetskaya LA, Berestenko AG
(2016) Photoelectronics 25:56
80. Glushkov AV (2008) Relativistic quantum theory. Quantum mechanics of atomic systems,
Astroprint, Odessa, 700p
81. Ivanov LN, Letokhov VS (1985) Com Mod Phys D 4:169; Ivanov LN, Ivanova EP,
Aglitsky EV (1988) Phys Rep 166:315
82. Ivanova EP, Ivanov LN, Glushkov AV, Kramida AE (1985) Phys Scripta 32:513
83. Glushkov AV, Kondratenko PA, Buyadgi VV, Kvasikova AS, Sakun TN, Shakhman AN
(2014) J Phys: Conf Ser 548:012025
84. Svinarenko AA, Glushkov AV, Khetselius OYu, Ternovsky VB, Dubrovskaya YuV,
Kuznetsova AA, Buyadzhi VV (2017) In: Rare Earth Element, (ed) Orjuela JEA. InTech,
pp 83–104
85. Glushkov AV, Khetselius OYu, Svinarenko AA, Buyadzhi VV, Ternovsky VB,
Kuznetsova AA, Bashkarev PG (2017) In: Uzunov DI (ed) Recent studies in perturbation
theory. InTech, pp 131–150
86. Baldwin GG, Salem JC, Goldansky VI (1981) Rev Mod Phys 53:687; Goldansky VI,
Letokhov VS (1974) JETP 67:513; Ivanov LN, Letokhov VS (1975) JETP 68:1748
87. Glushkov AV, Khetselius O, Gurnitskaya E, Loboda A, Sukharev D (2009) AIP Conf Proc
1102 (1):168
88. Ivanov LN, Letokhov VS, Glushkov AV (1991) Preprint of Inst. for Spectroscopy of USSR
Acad Sci (ISAN) AS-N5
89. Glushkov AV (2013) Advances in quantum methods and applications in chemistry, physics
and biology. In: Hotokka M, Brändas E, Maruani J, Delgado-Barrio G (eds) Progress in
theoretical chemistry and physics, vol 27. Springer, Cham, pp 161–177
90. Glushkov AV, Ambrosov SV, Loboda AV, Gurnitskaya EP, Khetselius OYu (2006)
Recent advances in theoretical physics and chemistry systems. In: Julien J-P, Maruani J,
Mayou D, Wilson S, Delgado-Barrio G (eds) Progress in theoretical chemistry and physics,
vol 15. Springer, Dordrecht, pp 285–299
91. Glushkov AV (2012) Quantum systems in chemistry and physics. In: Nishikawa K,
Maruani J, Brändas E, Delgado-Barrio G, Piecuch P (eds) Progress in theoretical chemistry
and physics, vol 26. Springer, Dordrecht, pp 231–252
92. Khetselius OYu, Zaichko PA, Smirnov AV, Buyadzhi VV, Ternovsky VB, Florko TA,
Mansarliysky VF (2017) Quantum systems in physics, Chemistry and biology. In: Tadjer A,
Pavlov R, Maruani J, Brändas J, Delgado-Barrio G (eds) Progress in theoretical chemistry and
physics, vol 30. Springer, Cham, pp 271–281
Part II
Molecular Structure and Dynamics
Difference of Chirality of the Electron
Between Enantiomers of H𝟐 X𝟐
1 Introduction
Only one form of enantiomeric pair is found in living systems. While both enan-
tiomers of sugar and amino acids are produced in the same quantity in laborato-
ries, only (D)-sugars and (L)-amino acids can be found in the systems. The mecha-
nism how this bias is generated is a long-standing puzzle [1]. On the origin of this
biomolecular homochirality, many hypotheses are proposed, and those are classi-
fied by some features, (1) terrestrial and extra-terrestrial, (2) biotic and abiotic, and
(3) probabilistic and deterministic. We do not know the solution, and however many
researchers believe that the realization of homochirality is deeply related to the par-
ity violation of Nature, which is given by the weak interaction of the standard model
of particle physics. The weak interaction is known to induce nuclear 𝛽 decay. Gauge
theory describes this interaction as well as electromagnetic interaction. Electromag-
netic interaction is formulated as U(1) gauge theory and is mediated by the photon.
On the other hand, the weak interaction is formulated as SU(2) gauge theory, and W
and Z bosons are mediator of this interaction. These bosons are coupled to ordinary
fermions, such as electrons, only by the V-A coupling, where V and A is the vec-
tor and axial-vector currents, respectively. Hence parity is violated maximally in the
weak interaction.
It is known that this parity violation makes the energy difference between enan-
tiomers. This energy difference between them, which is called parity-violating energy
shift, is very small [2]. In spite of the smallness, this energy is studied in many
works by computational methods. The parity-violating energy shift is dominantly
induced by virtual Z boson exchanges between nuclei and electrons. Virtual Z boson
exchanges is quantum effect in relation to uncertainty principle. Since Z boson is
very heavy as 90 GeV∕c2 , where c is the speed of light, this particle cannot be pro-
duced energetically, and, however, uncertainty relation allows this particle to affect in
a very restricted region. Single W boson exchange does not contribute to the parity-
violating energy, since if occurs, it is 𝛽 decay. The exchanges between electrons is
known to be subdominant for particularly heavy nuclei [3]. Since nuclei are localized
strongly, the existence of parity-violating energy means the existence of the nonzero
chirality density around nuclei. It is surprising that low energy electrons have polar-
ized chirality, since the electron mass, that is the interaction with Higgs field in the
vacuum, vanishes chirality for free electrons.
The parity-violating energy shift is studied by many researchers as one of solu-
tions for biomolecular homochirality. Due to this energy difference, the amount of
one of enantiomers may be slightly larger than the other through an enhancement
process, such as crystallization. In other viewpoints, some of the authors are inter-
ested in the total integrated chirality density of the electron in a molecule, whose
existence has already been reported [4]. If nonzero integrated chirality density of
the electron exists in enantiomers, its interaction rates through weak interaction are
different between enantiomeric pair. This difference of the reaction rate probably
generates the difference of the number density between enantiomers, which is indis-
pensable for the problem of the biomolecular homochirality. Even though this differ-
ence is not enough, we are interested in the distribution of the chirality density in a
molecule, since the chirality density is known to be proportional to the zeta potential
[5]. The zeta potential is the potential for the zeta force, which is the counter force to
the spin torque, defined only in quantum field theory [6, 7]. Hence chiral molecules
have nonzero zeta force distribution.
In this work, we study the integrated chirality density of the electron in H2 X2
molecules (X = O, S, Se, Te). This structure is one of the simplest chiral molecules
and is chosen in many works [4, 8–10]. The total chirality of the electron of this
structure is reported for H2 Te2 in Reference [4]. In this work, the chirality is investi-
gated also for H2 O2 , H2 S2 , and H2 Se2 in relation to the parity-violating energy shift
and the zeta potential.
Difference of Chirality of the Electron Between Enantiomers of H2 X2 97
2 Theory
First we briefly review the parity-violating energy. The dominant contribution to the
parity-violating energy is the parity odd interaction between electrons and nuclei.
This interaction is given as the coupling of vector and axial-vector currents of elec-
trons and nucleons. For low energy nuclei, nonrelativistic approximation is good,
and then the space-like component of vector current and the time-like component of
axial-vector current vanishes. The internal structure of nuclei is not affected by the
geometry of molecules, and hence the space-like component of axial-vector current
is considered to be negligibly small. Hence, the most important contribution arises
from the coupling of the time-like components of nucleon vector current and electron
axial-vector current. The Hamiltonian of this interaction is given as
∑ GF †
†
HPV = √ QW 𝜓̂ e 𝛾5 𝜓̂ e 𝜓̂ Nn 𝜓̂ Nn
n
(1)
n 2 2
where index n means the species of nuclei, GF = 1.166378 × 10−5 GeV−2 is Fermi
coupling constant [11], 𝜓̂ e and 𝜓̂ Nn are the field operators of electrons and nuclei,
and 𝛾5 ≡ i𝛾 0 𝛾 1 𝛾 2 𝛾 3 . The weak charge of a nucleus QnW is given as QnW = Z n (1 −
4 sin2 𝜃W ) − N n , where Z n and N n are the number of protons and neutrons in a nucleus
n and 𝜃W is the weak-mixing angle, sin2 𝜃W = 0.2313 [11]. The parity-violating
energy is calculated with state vector, |Ψ⟩,
and the parity-violating energy shift is defined as the energy difference between
enantiomers,
The parity-violating energy can be divided into contributions from each nucleus,
( )
G ∑ n G ∑ n n
EPV = √F QW d3 ⃗r⟨Ψ|𝜓̂ e† 𝛾5 𝜓̂ e 𝜓̂ N† 𝜓̂ Nn |Ψ⟩ = √F QW MPV . (4)
2 2 n ∫ n
2 2 n
n
This MPV is often used for parameterizing the contribution from each nucleus. The
density of nuclei is strongly localized, and hence nucleus density, 𝜓̂ N† 𝜓̂ Nn , can be
n
ℏc [ † ] ℏc ( † )
𝜙̂ 5 (x) = 𝜓̂ e (x)𝛾5 𝜓̂ e (x) = †
𝜓̂ eR (x)𝜓̂ eR (x) − 𝜓̂ eL (x)𝜓̂ eL (x) , (5)
2 2
where 𝜓̂ eL (x) ≡ 12 (1 − 𝛾5 )𝜓̂ e and 𝜓̂ eR (x) ≡ 12 (1 + 𝛾5 )𝜓̂ e are fields with the left-handed
and right-handed chirality, respectively. The zeta force density operator is defined
with the zeta potential, 𝜙̂ 5 , as
The zeta force is one of the torque in the equation of motion of the electron spin
defined in quantum field theory. The equation of motion of the spin is derived from
the time-derivative of the spin angular momentum. In quantum field theory, the spin
angular momentum density operator is represented as
1 †
ŝ ke (x) = ℏ𝜓̂ (x)Σk 𝜓̂ e (x), (7)
2 e
where Σk is the Pauli matrix in the four-component representation. The torque den-
sity for the electron spin is derived by the time-derivative of this operator, and the
time-derivative can be reduced by using Dirac equation.
gauge field operator. As a result, we obtain the equation of motion of the spin,
𝜕̂ske (x)
= ̂tek (x) + 𝜁̂ek (x), (9)
𝜕t
where the first term, ̂tek (x), is the spin torque density. The spin torque term is defined
so that this term matches the well-known spin torque term in relativistic quantum
mechanics. The Heisenberg equation of the spin angular momentum in quantum
mechanics is given by d⃗ŝ e ∕dt = −c𝜋⃗̂ × 𝛼⃗ [12]. The spin torque density operator is
defined with the relativistic stress tensor density, 𝜏̂eΠln (x), as
where 𝜖lnk is the Levi-Civita tensor. The relativistic stress tensor operator is given by
[5],
[ ( ) ]
𝜏̂eΠln (x) =
iℏc † ̂ l (x)𝜓̂ e (x) − D
𝜓̂ e (x)𝛾 0 𝛾 n D ̂ l (x)𝜓̂ e (x) † 𝛾 0 𝛾 n 𝜓̂ e (x) . (11)
2
Difference of Chirality of the Electron Between Enantiomers of H2 X2 99
The stress tensor is known to classify the chemical bond [13, 14]. In the following,
we call only ̂tek (x) the spin torque, and the sum of the terms in the right-hand side is
called the torque for the spin. Our equation of motion for the spin is recently shown
to be derived from the spin vorticity principle in a sophisticated way [15].
One may wonder whether this new contribution disturbs the consistency between
experimental observations and a prediction by quantum mechanics. The expecta-
tion value of the zeta force is zero after the integration over the whole space, since
the zeta force density operator is given as the gradient form of the zeta potential.
Hence, contributions from the zeta force are considered to be negligible in past exper-
iments. However, the contribution from the zeta force give a nonzero effect in a local
region, even after the integration over a restricted local region. Hence, the zeta force
is observable quantity if an experimental setup is carefully designed for this purpose.
In addition, the equation of motion from quantum field theory has another advan-
tage over that from quantum mechanics. In a time-independent stationary state of
the electron spin. the spin torque and zeta force are canceled out with each other and
the torque for the spin is zero at any point. Hence in quantum field theory a local
picture of the spin dynamics can be correctly described. In quantum mechanics, any
local spin dynamics cannot be predicted. The Heisenberg equation of the spin gives
generically nonzero torque for a local region even for a spin stationary state. This
is because the expectation value of quantum mechanics is defined as the integration
over the whole space and hence local description is theoretically out of scope.
3 Computational Details
Our equation is defined in quantum field theory, and hence a state should also be
prepared in the theory. However, a generic state based on quantum field theory is
not available for our purpose, since most computation code is based on quantum
mechanics. Hence in this work we use wave functions derived from ordinary elec-
tronic structure computations as a substitution. With the usage of these wave func-
tions, computations of physical quantities, parity-violating energy, zeta potential,
and so on, are performed by QEDynamics program package [16–18].
For ordinary electronic structure computations, we use DIRAC14 program pack-
age [19]. Four-component wave function by relativistic quantum mechanics can be
computed by this code, which is indispensable for the study of spin. The dyall.ae2z
basis set [20] is used for large components of all atoms. The small component
basis set is generated by restricted kinetic balance in the code. The effect of three-
component vector potential is ignored in our calculations, since it is quantitatively
small for states by quantum mechanics computations [21]. The structure of H2 X2
molecules are determined as follows. First, the geometrical optimization computa-
tions are performed by Hartree-Fock computations with Dirac-Coulomb Hamilto-
nian. The internuclear lengths between heavy atoms X, 1.390 Å for oxygen atoms,
2.058 Å for sulfur atoms, 2.333 Å for selenium atoms, and 2.729 Å for tellurium
atoms, while the internuclear lengths between X and H atoms, 0.9439 Å for oxygen
100 M. Senami et al.
Fig. 1 Geometry of H2 X2
molecule and the definition
of the dihedral angle
atom, 1.332 Å for sulfur atom, 1.455 Å for selenium atom, and 1.649 Å for tellurium
atom. In the following, these internuclear lengths are adopted for all dihedral angles.
For wave functions derived by computations as above, parity-violating energy,
n
MPV , total chirality, spin torque density, zeta force, and zeta potential are computed
with special attention to dihedral angle of H2 X2 . Since H2 X2 is a chiral molecule,
two choices of dihedral angle exist. Our definition of the dihedral angle is shown in
Fig. 1, which is the opposite to Reference [8, 9]. Results can easily be compared by
replacing the angle 360◦ − 𝜙. Note that parity-violating energy is known to be heav-
ily dependent on computational methods, geometry of molecules, basis sets and so
on. Our basis set is smaller than previous works, and hence there are some differences
between our results and previous works [4, 8–10] as discussed later.
For the purpose of the check of our electronic structure, our results of parity-violating
energy and contributions from heavy atoms are compared to previous works. In
X
Fig. 2, the contributions from heavy atoms to parity-violating energy, MPV , are shown
as a function of the dihedral angle for H2 X2 (X = O, S, Se, Te) molecules. All
X
molecules have similar pattern and it is seen that heavier X atoms give larger MPV
due to larger relativistic effects. The tendencies of these curves are consistent with
previous works [4, 8–10] qualitatively. The contribution of hydrogen atoms to EPV
is known to be much smaller than that of heavier atoms. Our values have some devi-
ation from previous works. This deviation is speculated to be due to the smallness
of our basis set. Parity-violating energy of H2 X2 and contributions from heavy atom
X are summarized in Table 1. The dihedral angle is chosen to be 45◦ or −45◦ as
the value reported in references, which have the same value with opposite sign. The
abbreviation, HF, means Hartree-Fock with Dirac-Coulomb Hamiltonian, CCSD is
coupled-cluster singles-and-doubles, and CISD is configuration interaction includ-
ing single and double excitations. As seen from this table, larger basis sets as triple
or quadruple zeta function are critically important, and results of double zeta basis
sets are much smaller than that of larger ones. Larger basis set is speculated to be
Difference of Chirality of the Electron Between Enantiomers of H2 X2 101
-4
1.0×10
4.0×10-6
5.0×10-5
2.0×10-6
MPVO [a.u.]
MPVS [a.u.]
0.0×100 0.0×100
-2.0×10-6 -5
-5.0×10
-6
-4.0×10
-4
-1.0×10
0 50 100 150 200 250 300 350 0 50 100 150 200 250 300 350
dihedral angle [degrees] dihedral angle [degrees]
(a) H2 O2 (b) H2 S2
4.0×10-3 4.0×10-2
2.0×10-3 2.0×10-2
[a.u.]
[a.u.]
0.0×100 0.0×100
Se
Te
MPV
MPV
-3 -2
-2.0×10 -2.0×10
-4.0×10-3 -4.0×10-2
0 50 100 150 200 250 300 350 0 50 100 150 200 250 300 350
required for the description of the cusp structure near nuclei. Since our triple zeta
result is well consistent with other triple zeta results, most deviation of our results
from others comes from the smallness of the basis set. In addition, the effect of post
Hartree-Fock computation is seen to be about 10%. Due to the limit of computational
resources, the computations of the latter part of this work is restricted to the basis
set, dyall.ae2z. Nevertheless, our wave functions are seen to be reasonable within
our computational methods from this table.
In addition, we investigate the spin torque and the zeta force of H2 O2 . Figure 3
shows the distributions of the spin torque, the zeta force, and their sum for H2 O2
with 𝜙 = 45◦ . It can be seen that the sum of the spin torque and the zeta force is
much smaller than the spin torque and the zeta force itself in the whole region. This
result is consistent with the fact that the nonzero spin torque is in balance with the
zeta force for the spin stationary state. Hence, although our computational result is
derived from wave functions of quantum mechanics, it is considered that we can use
these wave functions. The values of the norm of the spin torque and zeta force in the
vicinity of oxygen nuclei amount to 10−3 [a.u.], which is much large than that in the
vicinity of hydrogen nuclei. The smallness of the spin torque around hydrogen nuclei
102 M. Senami et al.
Table 1 Parity-violating energy of H2 X2 and contributions from heavy atom X are summarized.
The dihedral angle is chosen to be 45◦ or −45◦ , which have the same value with opposite sign. HF
means Hartree-Fock with Dirac-Coulomb Hamiltonian
Molecule Method Basis set for X |MPVX
| [a.u.] |EPV | [a.u.] Reference
H2 O2 HF dyall.ae2z 3.3401 × 10−6 3.8853 × 10−19 This study
dyall.ae3z 5.1839 × 10−6 6.0300 × 10−19 This study
dyall.ae3z − 6.051 × 10−19 [10]
dyall.ae4z − 6.376 × 10−19 [10]
cc-pVDZ+3p 5.801 × 10−6 − [9]
[22]
25s25p5d [8] 6.057 × 10−6 − [9]
aug-cc-pVDZ 3.729 × 10−6 − [8]
[22]
aug-cc-pVTZ 4.239 × 10−6 − [8]
[22]
aug-cc-pVQZ 4.687 × 10−6 − [8]
[22]
25s25p5d 6.057 × 10−6 − [8]
CISD cc-pVDZ+3p 5.410 × 10−6 − [9]
CCSD dyall.ae3z − 5.323 × 10−19 [10]
dyall.ae4z − 5.583 × 10−19 [10]
cc-pVDZ+3p 5.299 × 10−6 − [9]
H2 S2 HF dyall.ae2z 7.0563 × 10−5 1.6416 × 10−17 This study
cc-pCVTZ [22] − 1.825826 × [10]
10−17
cc-pVDZ+3p 8.916 × 10−5 − [9]
25s25p5d 9.581 × 10−5 − [8]
CCSD cc-pCVTZ [22] − 1.82103 × 10−17 [10]
cc-pVDZ+3p 9.283 × 10−5 − [9]
H2 Se2 HF dyall.ae2z 2.3917 × 10−3 1.6334 × 10−15 This study
25s25p5d 3.586 × 10−3 − [8]
CCSD dyall.cv3z [23] − 2.115 × 10−15 [10]
H2 Te2 HF dyall.ae2z 2.3842 × 10−2 2.7769 × 10−14 This study
25s25p5d 3.149 × 10−2 − [8]
CCSD dyall.cv3z − 3.289 × 10−14 [10]
O
is consistent with our previous results [6]. The nonzero MPV means the existence of
the zeta force around oxygen nuclei, since the chirality density is proportional to the
zeta potential and the zeta potential is the potential for the zeta force. Hence this
result is consistent with the existence of the parity-violating energy.
In Fig. 4, the integrated chirality density as a function of the dihedral angle is
shown for (a) H2 O2 , (b) H2 S2 , (c) H2 Se2 , and (d) H2 Te2 molecules. Our result of
Difference of Chirality of the Electron Between Enantiomers of H2 X2 103
Fig. 3 The distributions of the magnitude and the direction of a the spin torque distribution,
b the zeta force, and c the sum of the spin torque and zeta force are shown for H2 O2 molecule.
The dihedral angle is 45◦
H2 Te2 is consistent with the result reported in Reference [4], and it is confirmed that
the electron chirality is nonzero in chiral molecules. The integrated chirality density
of H2 Te2 has almost the same dependence on the dihedral angle. However, the inte-
grated chirality density of H2 O2 and H2 S2 are almost opposite to the parity-violating
X
energy, which is determined dominantly from MPV shown in Fig. 2. Moreover, the
integrated chirality density of H2 Se2 has different oscillation pattern from the parity-
violating energy. Although we guessed that the integrated chirality density and the
parity-violating energy of H2 X2 have some correlation as in Reference [4], our inte-
grated chirality density is not inconsistent with the parity-violating energy, since the
parity-violating energy is determined dominantly only by the chirality density nearby
heavy nuclei. Nevertheless, we should improve our computations with larger basis
set and perform post Hartree-Fock computations in order to check our results.
In Fig. 5, the distributions of zeta potential around one Te atom of H2 Te2 at the
dihedral angle, (a) 15◦ , (b) 45◦ and (c) 90◦ , are shown on the xy-plane for the z coordi-
nate on Te atoms. Our results are well consistent with those reported in Reference [4].
We have shown only the results for 𝜙 = 15◦ , 45◦ , 90◦ , while at other dihedral angles,
our results are consistent with Reference [4]. The distribution pattern of zeta poten-
tial agrees with their results well. The difference of the values arises from the factor
ℏc∕2, which is the coefficient of the zeta potential over the chirality density. For
comparison, the same figure for H2 O2 is shown in Fig. 6 at the dihedral angle, (a)
15◦ , (b) 45◦ and (c) 90◦ . The results are shown on the xy-plane for the z coordinate
104 M. Senami et al.
1.0 ×10-7
4.0 ×10-7
chirality density [a.u.]
-7
-2.0× 10
-5.0 ×10-8
-4.0× 10-7
-1.0 ×10-7
0 50 100 150 200 250 300 350 0 50 100 150 200 250 300 350
dihedral angle [degrees] dihedral angle [degrees]
(a) H2 O2 (b) H2 S2
-6 -5
1.0 ×10 2.0 ×10
-5
1.5 ×10
Fig. 4 The integrated chirality density as a function of the dihedral angle for a H2 O2 , b H2 S2 ,
c H2 Se2 , and d H2 Te2 molecules
Fig. 5 The distribution of zeta potential of H2 Te2 at the dihedral angle, a 15◦ , b 45◦ and c 90◦ .
The result is shown on the xy-plane for the z coordinate on Te atoms
at one O atom. The localization of the innermost core electrons are strongly different
between O and Te atoms, and hence the distribution pattern of the zeta potential is
largely extended. The sign of the zeta potential at the position of a nucleus is the
same for 15–45◦ , and opposite for 90◦ , this corresponds to the dependence of MPV X
Fig. 6 The distribution of zeta potential of H2 O2 at the dihedral angle, a 15◦ , b 45◦ and c 90◦ . The
result is shown on the xy-plane for the z coordinate on O atoms
5 Conclusion
References
1. Meyerhenrich U (2008) Amino acids and the asymmetry of life. Springer, Heidelberg
2. Mason SF (1984) Nature 311:19
3. Hegstrom RA, Rein DW, Sandars PGH (1980) J Chem Phys 73:1
4. Bast R, Koers A, Gomes ASP, Iliaš M, Visscher L, Schwerdtfeger P, Saue T (2011) Phys Chem
Chem Phys 13:864
5. Tachibana A (2001) J Chem Phys 115:3497; J Mol Model 11:301 (2005); J Mol Struct:
(THEOCHEM), 943:138 (2010). see also Tachibana A (2017) New aspects of quantum elec-
trodynamics. Springer
106 M. Senami et al.
6. Fukuda M, Senami M, Tachibana A (2013) Prog Theor Chem Phys 27:131; Fukuda M, Soga
K, Senami M, Tachibana A (2016) Int J Quant Chem 116:920
7. Senami M, Nishikawa J, Hara T, Tachibana A (2010) J Phys Soc Jpn 79:084302; Hara T,
Senami M, Tachibana A (2012) Phys Lett A 376:1434; Fukuda M, Soga K, Senami M,
Tachibana A (2013) Phys Rev A 93:012518; Fukuda M, Ichikawa K, Senami M, Tachibana
A (2016) AIP Adv 6:025108
8. Laerdahrl JK, Schwerdtfeger P (1999) Phys Rev A 60:4439
9. Thyssen J, Laerdahrl JK, Schwerdtfeger P (2000) Phys Rev Lett 85:3105
10. Shee A, Visscher L, Saue T (2016) J Chem Phys 145:184107
11. Patrignani C et al (2016) Particle data group. Chin Phys C 40:100001
12. Sakurai JJ (1976) Advanced quantum mechanics. (Addison-Wesley, Reading, Mass)
13. Tachibana A (2004) Int J Quantum Chem 100:981
14. Ichikawa K, Nozaki H, Komazawa N, Tachibana A (2012) AIP Adv 2:042195
15. Tachibana A (2012) Math J Chem 50(669); In: (2013) Ghosh SK, Chattaraj PK (eds) Concepts
and methods in modern theoretical chemistry. CRC, Boca Raton, FL, pp 235–251
16. Senami M, Ichikawa K, Tachibana A, QEDynamics http://www.tachibana.kues.kyoto-u.ac.jp/
qed
17. Ichikawa K, Fukuda M, Tachibana A (2013) Int J Quantum Chem 113:190
18. Senami M, Miyazato T, Takada S, Ikeda Y, Tachibana A (2013) J Phys Conf Ser 454:012052;
Senami M, Ogiso Y, Miyazato T, Yoshino F, Ikeda Y, Tachibana A (2013) Trans Mat Res Soc
Jpn 38(4):535
19. Saue T, Visscher L, Jensen HJAa, Bast R, Bakken V, Dyall KG, Dubillard S, Ekström U, Eliav
E, Enevoldsen T, Faßhauer E, Fleig T, Fossgaard O, Gomes ASP, Helgaker T, Lærdahl JK, Lee
YS, Henriksson J, Iliaš M, Jacob ChR, Knecht S, Komorovský S, Kullie O, Larsen CV, Nataraj
HS, Norman, P Olejniczak G, Olsen J, Park YC, Pedersen JK, Pernpointner M, di Remigio R,
Ruud K, Sałek P, Schimmelpfennig B, Sikkema J, Thorvaldsen AJ, Thyssen J, van Stralen J,
Villaume S, Visser O, Winther T, Yamamoto S (2014) DIRAC a relativistic ab initio electronic
structure program. http://www.diracprogram.org
20. Dyall KG (2012) Theor Chem Acc 99:366 (1998); Theor Chem Acc 115:441 (2006); Theor
Chem Acc 117:483 (2007); J Phys Chem A 113:12638 (2009); Theor Chem Acc 131:1217
21. Senami M, Ikeda Y, Fukushima A, Tachibana A (2010) Jpn J Appl Phys 49:115002
22. Dunning TH Jr (1989) J Chem Phys 90:1007; Kendall RA, Dunning TH Jr, Harrison RJ, J
Chem Phys 96:6796 (1992); Woon DE, Dunning TH Jr, J Chem Phys 98:1358 (1993)
23. Dyall KG (2009) J Phys Chem A 113:12638; Theor Chem Acc 108:335 (2002); Chem Acc
115:441 (2006); Theor Chem Acc 117:483 (2007)
A Crystallographic Review of Alkali
Borate Salts and Ab Initio Study
of Borate Ions/Molecules
Cory C. Pye
Abstract The crystal structures of alkali metal borate salts are reviewed. A wide
diversity of structures is noted. Structures with discrete ions containing two or more
boron atoms are targeted for further study with ab initio methods (HF, B3LYP, and
MP2) using modest basis sets (6-31G*, 6-31+G* and 6-311+G*). The ions iden-
tified for study are: [B3O6]3−, [B3O5(OH)2]3−, [B3O4(OH)4]3−, [B3O3(OH)4]−,
[B4O5(OH)4]2−, [B5O6(OH)4]−, [B2O5]4−, and [B4O9]6−. Some structurally related
ions are examined, and an investigation of the diborates launched, including [B2O
(OH)6]2−, observed in the magnesium salt, and [B2(OH)7]−, postulated as the
intermediate responsible for signal exchange between borate anion and boric acid in
11
B NMR. The B-O bond distances and general structures are in good agreement
with both crystallographic data and previous ab initio calculations.
1 Introduction
In the preceding paper, the importance of the speciation of boron (III) to geo-
chemistry, oceanography, and the nuclear industry, was discussed [1]. In addition,
the crystal structures of boron oxide, boric acid, and monomeric forms of sodium
and lithium borates were presented and discussed. Molecular forms exist for the
orthoboric acid [B(OH)3] and metaboric acid [B3O3(OH)3], and certain lithium and
sodium borates contain discrete [B(OH)4]−, [BO2(OH)]2−, and [BO3]3− ions.
The structure, energy, and vibrational frequencies of the crystallographically
C. C. Pye (✉)
Department of Chemistry, Saint Mary’s University, Halifax, NS B3H 3C3, Canada
e-mail: [email protected]
96.02
3 5 4 423 Pbca 8.804,18.371,10.924 1.425–1.531 1.351–1.398 Layers of B5O6
polyhedra [10]
3 5 2 523 Pca21 11.2373,6.0441,11.1336 291 1.440–1.510 1.367–1.377 Layers of B5O6
polyhedra [11]
1 2 10 Borax (sodium C2/c 11.89, 10.74, 12.19 [12]
tetraborate decahydrate)
C2/c 11.858,10.674,12.197 298 1.46–1.54 1.32–1.40 Chains of [B4O5(OH)4]2−
106.68 ions [13]
298 C2/c 11.885,10.654,12.206 297 1.464–1.500 1.363–1.374 Neutron diffraction [14]
(continued)
109
Table 1 (continued)
110
Stoichiometry Name Synth T(K) Space group Unit cell X-ray T(K) B-O (Å) Comments
NalBmHnO(l+3m+n)/ (Å or deg.) ranges
2
(continued)
Table 2 (continued)
114
Stoichiometry Name Synth T(K) Space group Unit cell (Å or deg.) X-ray T(K) B-O (Å) ranges Comments
KlBmHnO(l+3m+n)/
2
Stoichiometry Name Synth T(K) Space group Unit cell (Å or deg.) X-ray T(K) B-O (Å) ranges Comments
LilBmHnO(l+3m+n)/
2
initially to result in partial hydrolysis into discrete structural motifs already present
in the solid state, followed by equilibration in solution.
Ab initio calculations have been carried out on some of the discrete ions seen in
the crystal structures. Gupta and Tossell fixed the symmetry of B2O(OH)4 to C2v
(B-O fixed at 1.353 Å (HF/STO-3G) and 1.375 Å (4-31G), and found that the B-O-B
linkage was bent [94]. They also found B-O distances of 1.430 Å and 1.279 Å for
B3O63− at HF/STO-3G, which compares well with the crystal structure. Zhang et al.
[95] completely optimized B2O(OH)4 with symmetry C2v (HF/STO-3G, 6-31G*), Cs
(HF/6-31G) and C2 (HF/STO-3G, 3-21G*, 4-31G, 6-31G, 6-31G*). They also
calculated the B2O(OH)2− 6 (C2v), B2O(OH)− −
5 (Cs), B3O3(OH)4 (C2v) and B3O3
(OH)5 (Cs) ions at HF/STO-3G. Oi calculated B2O(OH)4 (C2), B2O(OH)−
2−
5 (C1), and
−
B2O(OH)2− 6 (C2) at HF/6-31G* [96]. In addition, Oi also calculated B3O3(OH)4
2− 2− −
(C2), B3O3(OH)5 (C1), B4O5(OH)4 (C2), and B5O6(OH)4 (S4) at HF/6-31G* [97].
A combined Raman and DFT (B3LYP/aug-cc-pVDZ) investigation of B2O(OH)4,
B2O(OH)− 2− − 2− 3−
5 , B2O(OH)6 , B3O3(OH)4 , B3O3(OH)5 , B3O3(OH)6 , B4O5(OH)4 ,
2−
−
and B5O6(OH)4 was presented by Zhou et al [98]. In addition to these ions, two
heptamers B7O9(OH)2− 5 were calculated by Beckett et al. at B3LYP/6-311++G(d, p)
[99].
2 Methods
Calculations were performed using Gaussian 03 [100]. The MP2 calculations use
the frozen core approximation. The geometries were optimized using a stepping
stone approach, in which geometries at the levels HF/6-31G*, HF/6-31+G*, HF/
6-311+G*, B3LYP/6-31G*, B3LYP/6-31+G*, B3LYP/6-311+G*, MP2/6-31G*,
MP2/6-31+G* and MP2/6-311+G* were sequentially optimized, with the geom-
etry and molecular orbitals reused for the subsequent level. Default optimization
specifications were normally used. After each level, where possible, a frequency
calculation was performed at the same level and the resulting Hessian was used in
the following optimization. Z-matrix coordinates constrained to the appropriate
symmetry were used as required to speed up the optimizations. Because frequency
calculations are done at each level, any problems with the Z-matrix coordinates
would manifest themselves by giving imaginary frequencies corresponding to
modes orthogonal to the spanned Z-matrix space. The Hessian was evaluated at the
first geometry (Opt = CalcFC) for the first level in a series in order to aid geometry
convergence.
120 C. C. Pye
In our previous work, the structure, energy, and vibrational spectra of orthoboric
acid, metaboric acid, and tetrahydroxoborate was thoroughly discussed. We now
focus on the polyborate ions. First we discuss those observed in the crystal structure
of alkali metal borates, followed by structurally related ions. The more
highly-charged ions might be observed in low-water, high ionic-strength environ-
ments. We also discuss other possibilities.
3.1.1 [B3O3(OH)4]−
3.1.2 [B3O3(OH)5]2−
The structure of the triborate ion [B3O3(OH)5]2− is given in Fig. 1. Neither of the
two Cs forms is stable and both desymmetrize to the corresponding C1 forms. Of
these, structure #2 is 2.1–3.4 kJ/mol more stable than structure #1.
3.1.3 [B3O3(OH)6]3−
The structure of the triborate ion [B3O3(OH)6]3− is given in Fig. 1. First, two D3h
structures were tried. Neither were minima, and the number of imaginary
A Crystallographic Review of Alkali Borate Salts … 121
C2 #1 C2 #2 C2 #3/C1 #4 C2 #4/C1 #2 Cs #2
C1 #5 C1 #6 C1 #7 Cs #3
Cs #1 Cs #2 C1 #1 C1 #2
Cs #1 Cs #2 C3 #1 C3 #2 C2v #2
C2 #1 C1 #1 C1 #2 C1 #3
frequencies suggested many possible structures of lower symmetry. The two pos-
sible D3 structures (via A1”) coalesced into one, but it was not a minimum. The two
possible C3h structures (via A2’) coalesced into one, and it too is not a minimum.
While desymmetrization along the E” and E’ modes should lead to a C2 and Cs
structure, respectively, it was thought that it would be advantageous to first
desymmetrize to the corresponding C2v structures (which might ascend in symmetry
back to the corresponding D3h structures). Neither of these was a minimum, as
expected, and C2v #1 did indeed ascend to D3h #1 at the HF levels. The D3 and C3h
forms desymmetrize into the stable C3 #1 and #2 forms, respectively, along the A2
and A” modes, respectively. The two possible C2v structures could desymmetrize
into either two C2 structures, or into one of four possible Cs structures. The two C2
structures coalesce into a structure that is only stable at MP2/6-31G*. Cs #3 (from
C2v #1) and Cs #4 (from C2v #2) ascend in symmetry to either C3h or C2v #1.
Neither of the Cs structures are stable. The C2 structure desymmetrizes into C1 #1 at
MP2/6-31+G* and MP2/6-311+G*, and ascends in symmetry to C3 #1 at the HF
and B3LYP levels. The Cs structures desymmetrize to the stable C1 #2 and 3 at
most levels. The C3 #1 structure is the most stable.
3.1.4 [B3O4(OH)4]3−
The structure of the triborate ion [B3O4(OH)4]3− is given in Fig. 2. The C2v
structure has at least three imaginary frequencies of irreducible representation A2,
B1 and B2. These suggest desymmetrization to a C2 or to two different Cs structures,
respectively. None of these are energy minima (except Cs #1 at B3LYP/6-31G*)
and all desymmetrize to C1 structures #1–3, respectively. C1 #1 is the most stable.
This ion has been characterized crystallographically as the sodium and rubidium
salt (see Tables 1 and 4).
3.1.5 [B3O5(OH)2]3−
The structure of the triborate ion [B3O5(OH)2]3− is given in Fig. 2. The two C2v
structures are unstable and all contain an imaginary A2 mode, leading to two C2
structures that coalesce into the most stable minimum. The higher energy C2v #2
structure also contains a B2 mode. The lower energy C2v #1 structure also contains
either a B2 mode (B3LYP/6-31G*) or B1 mode (MP2/6-31+G* and MP2/6-311
+G*). The Cs #1 structure, derived from C2v #2, is only stable at B3LYP/6-31G*.
Cs #2, derived from C2v #1, coalesces into Cs #1. Cs #3, also derived from C2v #1, is
not stable. The C1 #1 structure, derived from Cs #1, coalesces into C2, except at the
stable B3LYP/6-31+G* and B3LYP/6-311+G* levels, whereas the C1 #2 structure,
derived from Cs #3, coalesces into C2 at both levels. This ion has been characterized
as the sodium salt (see Table 1).
A Crystallographic Review of Alkali Borate Salts … 123
C2v C2 Cs #1 Cs #2 C2v #1
D3h C1 #1 C1 #2 C1 #3 C2v #2
C2 Cs #1 Cs #3 C1 #1
3.1.6 [B3O6]3−
The only structure of the triborate ion [B3O6]3−, of D3h symmetry, is given in Fig. 2.
This ion is characterized crystallographically in sodium, potassium, rubidium, and
cesium metaborate (see Tables 1, 2, 3, 4 and 5). There does not appear to be any
systematic trend in the range of the B-O distances upon varying the alkali metal.
3.2.1 [B4O5(OH)4]2−
The structure of the tetraborate ion, [B4O5(OH)4]2−, is given in Fig. 3. Initially, four
structures of C2v symmetry were tried. None of these were energy minima. All of
these had both an A2 and a B2 imaginary mode, and, at some levels, some have a B1
imaginary mode. The four structures of C2 symmetry derived by desymmetrizing
along the A2 mode were all minima, but only three unique structures were found, as
some coalesced. The four structures of Cs symmetry derived by desymmetrizing
along the B2 mode (#1–#4) coalesced into two different structures, but only Cs #2
was a minima, at all levels except B3LYP/6-31G* and MP2/6-31G*. The four
structures of Cs symmetry derived by desymmetrizing along the B1 mode (#5–#8)
did not give minimum energy structures, and Cs #8 ascended in symmetry to C2v
124 C. C. Pye
#3. Cs #1 desymmetrized to the stable C1 #1. The C2 #2, C2 #3, and C1 #1 are quite
close in energy (5 kJ/mol). This ion has been characterized as a salt of all the alkali
metal ions (except lithium), and the sodium salt is commonly known as borax.
An alternative structure might exist where two of the hydroxyls connect to the
non-bridgehead boron atoms. Initially, four structures of C2v symmetry (#5–#8)
were tried. One of these (C2v #6) was an energy minima at all levels except B3LYP/
6-31G*. For this structure at this level, and for the other three structures, there exists
A2 and B1 imaginary modes. This suggests desymmetrization to C2 #5–#8 and Cs
#9–#12, respectively. The C2 #6 (B3LYP/6-31G* only) and #7 structures are stable.
All other attempts at structures of C2 symmetry revert to either C2 #7 or C2v #6. The
Cs #10 (B3LYP/6-31G*) and Cs #11 is stable. All other attempts at structures of Cs
symmetry revert to either Cs #11 or C2v #6.
3.2.2 [B4O9]6−
Another tetraborate ion, observed in the lithium salt (Table 3), is shown in Fig. 3.
Initially we considered structures of C2h and C2v symmetry, of which there are two
each. None of these is an energy minimum. For C2h #1, there are imaginary modes
of irreducible representation Au (desymmetrization to C2), Bg (desymmetrization to
Ci), and, at MP2/6-311+G*, Bu (desymmetrization to Cs #2). For C2h #2, there are
Au imaginary modes, and at MP2/6-31+G* and MP2/6-311+G*, Bu imaginary
modes (desymmetrization to Cs #3). For C2v #1, there are imaginary modes of
irreducible representation A2 (desymmetrization to C2) and B2 (desymmetrization to
Cs), whereas for C2v #2, there is only an A2 imaginary mode, giving rise to the
stable C2 #4. Desymmetrization of C2h #1 to a Ci structure results in ascent in
symmetry to C2h #2. Desymmetrization of C2v #1 to a Cs structure usually results in
ascent in symmetry to C2h #2, except for B3LYP/6-311+G* and MP2/6-311+G*,
where the Cs #1 structure remains. None of the three Cs structures are energy
minima. The Cs #1 structure desymmetrizes to the stable C1 #1, whereas attempts to
desymmetrize Cs #2 and Cs #3 result in ascent in symmetry to C2 #4.
3.3.1 [B5O6(OH)4]−
C2 #1/C1 #4 C2 #2 C2 #3 Cs #1 Cs #2/C1 #2
Cs #5 Cs #6 Cs #7 C1 #1 C2v #5
C2v/2 #6/
C2v #7 C2v #8 C2 #7 Cs #11
Cs #10
C2 #4 Cs #1 Cs #2 Cs #3 C1 #1
3.3.2 [B5O6(OH)6]3−
The pentaborate ion, [B5O6(OH)6]3−, is given in Fig. 4. We initially started with two
C2v structures and two C2 structures. Neither C2v structure is stable. Both have
imaginary A2, B1, and B2 modes at most levels. Both C2 structures are stable at most
levels, with the exception of B3LYP/6-311+G* (C2 #1) and B3LYP/6-31+G* (C2
#2), which desymmetrize to stable C1 #1 and #2, respectively. The C2 #3 and #4
structures, derived from C2v #1 and #2, are unstable at all levels and desymmetrize to
the stable C1 #3 and #4. At B3LYP/6-31G* and MP2/6-31G*, one of the central B-O
bonds in C1 #3 breaks. Attempts to locate the Cs #1 and #2 structures also result in
breaking of one of the central B-O bonds. The Cs #3 and #4 structures are not stable,
C2v #1 C2v #2 C2 #3 C2 #4 Cs #3
Cs #4 C1 #3 C1 #4 C1 #5 C1 #6
D2d S4 D2 Cs
C1 #1 C2 #2 C2 #3 C1 #2
3.3.3 [B5O6(OH)8]5−
3.4.1 [B2O5]4−
D2d C2 #1 Cs #1 D2h
C2v #1 C2 #1 C2 #2 Cs #1 Cs #2
C2v #2 C1 #1 C1 #2 C1 #3 C1 #4
tried. The D2d structure was stable at B3LYP/6-31G* and MP2/6-31G*, but had an
imaginary E mode at the other levels, suggesting desymmetrization to either a C2 #1
or Cs #1. The C2 #1 is stable at all levels except HF/6-31+G* and HF/6-31+G*, for
which the Cs #1 is stable. The D2h structure is unstable at all levels. Desym-
metrization along an imaginary Au mode at all levels results in ascension in sym-
metry to the D2d structure. Desymmetrization along the B2u, B3u, and B3g modes
give unstable C2v #1 (all levels except B3LYP/6-31G*), C2v #2 (MP2/6-31+G* and
MP2/6-311+G*), and C2h (B3LYP/6-31+G* and B3LYP/6-311+G*) structures.
Desymmetrization of these structures leads to structures already observed.
3.4.2 [B2(OH)7]−
In the polyborate structures discussed thus far, any oxygen that is bound to two
boron atoms is not bound to a hydrogen. However, the diborate species [B2(OH)7]−
with a bridging hydroxide was postulated as the intermediate to explain the boric
acid–borate interchange in aqueous solution as studied by 11B NMR [101]. Such an
intermediate may lie on the reaction path to condensation to form the other
polyborates. It is expected that the hydrogen of the bridging oxygen would be quite
acidic and easily lost.
The diborate ion, [B2(OH)7]−, is given in Fig. 5. Initially, two C2v structures
were tried. These had numerous imaginary A2, B1, and B2 frequencies, which
suggested desymmetrization to two C2 and four Cs structures. Neither C2 structure
was stable, possessing an imaginary B mode. The Cs #1 and Cs #2 structures,
derived from C2v #1, were not stable, containing A” imaginary frequencies. The Cs
#3 and Cs #4 structures, derived from C2v #2, either dissociated into a hydrogen
bonded [B(OH)4]−…B(OH)3 hydrogen bonded complex (Cs #3) or coalesced into
Cs #2. The C2 structures desymmetrized into the corresponding stable C1 #1 and #2
structures. The Cs structures desymmetrized into the corresponding stable C1 #3 and
#4 structures. The four stable structures are quite close in energy, and the gas-phase
association energy of boric acid and borate is in the range −47.9 to 108.5 kJ/mol.
3.4.3 [B2(OH)6]0
We were curious to see whether a dimer of boric acid could exist possessing a
single hydroxyl bridge. The bridge necessarily converts one of the boron atoms into
tetrahedral, and under the constraint of Cs symmetry, results in eight possible
structures. None of these are stable at the HF/6-31G* level, resulting in fragmen-
tation of the B-O bond to give hydrogen bonded structures. In addition, structure Cs
#1 is unstable at all levels investigated. These results suggest that such a structure
does not exist.
Next, a dimer of boric acid containing two bridging hydroxyls was investigated.
The optimized structures are shown in Fig. 6. Initially two structures of D2h
symmetry were considered. Neither of these structures were stable, possessing
A Crystallographic Review of Alkali Borate Salts … 129
Ci #1 C2 #1 Cs #10 C2 #5 Cs #9
C2v/2 #4 C2v #3 C1 #2
C2 #1 C2 #2 C2 #3 Cs/1 #3 Cs/1 #4
Cs #7 Cs #8 Cs #9 Cs/1 #5 Cs/1 #6
imaginary B3g and B2u modes at all levels, suggesting desymmetrization to the C2h
#1 and #3, and C2v #1 and #2 structures. The D2h #2 structure possessed an Au
mode at all levels (as did D2h #1 at MP2/6-311+G*), suggesting desymmetrization
to D2 #2 (D2 #1). The D2h structures also contained a B2g imaginary mode at some
levels suggesting desymmetrization (D2h #1: B3LYP/6-311+G* to C2h #2; D2h #2:
HF/6-311+G* to C2h #4).
To our surprise, the C2h #1 structure was stable at all B3LYP levels and at MP2/
6-31+G*, but possessed a Bg imaginary mode at the other levels, suggesting
130 C. C. Pye
desymmetrization to Ci #1. The C2h #3 structure was stable at all levels! The C2h #2
structure (B3LYP/6-311+G*) breaks apart into two boric acid molecules. The C2v
#1 structure has an A2 imaginary mode at all levels leading to C2 #1. The D2 #1
structure (MP2/6-311+G*) has an imaginary B3 and B2 mode, leading to C2 #2 and
C2 #3. The C2v #2 structure has an A2 and B1 imaginary mode at all levels leading to
C2 #4 and Cs #9. The D2 #2 structure has an imaginary B3 and B2 modes at all
levels, leading to C2 #5 and C2 #6. The C2h #4 structure (HF/6-311+G*) is
unstable, possessing both a Bg and Au mode, suggesting desymmetrization to Ci #2
and C2 #7.
The Ci #1 structure is stable at all levels except HF/6-31G* and HF/6-311+G*,
where it dissociated. The C2 #1 structure is only stable at HF/6-31+G*. It disso-
ciates at HF/6-31G*, HF/6-311+G*, and B3LYP/6-311+G*. At the other levels, it
ascends in symmetry via C1 #1 to the new Cs #10 structure, which exists at all
levels. The C2 #2 structure ascends in symmetry to C2h #1. The C2 #3 structure
coalesces to C2 #1. The C2 #4 structure ascends in symmetry to the new C2v #3. The
unstable Cs #9 structure only exists at MP2/6-31G*, ascending in symmetry to C2v
#3 otherwise. It desymmetrizes to the stable C1 #2 at MP2/6-31G*. The C2 #5
structure is stable at all levels, but the C2 #6 structure ascends in symmetry to C2v
#3. The Ci #2 and C2 #7 ascend in symmetry to C2h #3 and C2v #3, respectively.
To summarize these results, there are at least four stable structures at all levels.
Their energy ordering is as follows: C2v #3 (0.0 kJ/mol) < C2h #3 (5.7–9.8 kJ/mol)
< Cs #10 (8.1–13.1 kJ/mol) < C2 #5 (16.4–21.6 kJ/mol). The endothermic
gas-phase dimerization energy of boric acid lies in the range 35.8–103.8 kJ/mol.
3.4.4 [B2(OH)5]+
Another structure of interest is the cationic [B2(OH)5]+ (Fig. 6). Initially we tried
four structures of C2v symmetry. None of these was a minimum except for C2v #4 at
all levels except MP2/6-311+G*. All non-minima had imaginary A2 modes, and in
some cases, B1 modes. Deymmetrization along the A2 modes led to the stable C2
#1–#4 structures. Deymmetrization along the B1 modes led to the unstable Cs #7–
#9 structures. In addition, there are six other Cs structures obtainable by flipping the
hydrogen atoms. The Cs #1 and #3 structures are unstable at all levels. The Cs #2
structure is only stable at HF/6-31G*, HF/6-31+G*, and B3LYP/6-31+G*. The Cs
#4, #5, and #6 structures are stable at all levels except MP2/6-311+G*. These Cs
structures desymmetrize into the corresponding C1 #1–#9 structures. The C1 #1–#6
structures are stable, but the C1 #7–#9 structures convert to other structures already
obtained. Of these structures, Cs #4 is the most stable, followed by C2v #4 (8.0–
10.2 kJ/mol). The structure (with hydroxide) is thermodynamically unstable rela-
tive to two boric acid molecules (880–1020 kJ/mol).
A Crystallographic Review of Alkali Borate Salts … 131
3.4.5 [B2O(OH)6]2−
The diborate ion [B2O(OH)6]2− has been observed as the magnesium salt in the
mineral pinnoite. Stadler found that the space group of pinnoite was P42 or P42/m,
with a = 7.617(2) and c = 8.190(2), and proposed that the ion was [B2O(OH)6]2−
[102]. The space group was confirmed as P42 (a = 7.62(1), c = 8.19(1) by Paton
and MacDonald and the structure of the ion confirmed [103]. The structure was
further refined by Krogh-Moe [104]. Its structure is shown in Fig. 7. We initially
tried two C2v structures. Both were unstable and possessed A2 and B2 imaginary
modes. In addition C2v #1 possessed an imaginary B1 mode. The two C2 structures
thus derived were stable (with C2 #2 coalescing into C2 #1 at B3LYP/6-31G* and
MP2/6-31G*). None of the three Cs structures were stable, and desymmetrized to
the corresponding C1 structures. C1 #3 coalesced into C1 #1 at B3LYP/6-31G* and
MP2/6-31G*. Other possibilities include two Cs structures (#4 and #5), related to
the C2v structures by rotation of one of the hydroxyls. Neither of these are stable,
and desymmetrize to the corresponding C1 structures.
3.4.6 [B2O(OH)5]−
The diborate ion [B2O(OH)5]− was initially considered to contain a only a single
oxo bridge (Fig. 8). Eight such structures of Cs symmetry were considered, with the
BO3 unit in the plane of symmetry. None of these was stable and led to the
corresponding C1 structures. In some cases, these coalesced. It could also poten-
tially exist as a (μ-O)(μ-OH) doubly bridged dimer of C2v symmetry. This structure
is unstable, and has imaginary A2, B1, and B2 modes leading to C2, Cs #9, and Cs
#10 structures. In Cs #10, one of the B-O(H) bonds has broken. All of these
structures also have imaginary modes, leading potentially to C1 #9−#11,
C2v #1 C2 #1 C2 #2 C2v #2
Cs #1 Cs #2 Cs #3 Cs #4 Cs #5
C1 #1 C1 #2 C1 #3 C1 #4 C1 #5
Cs #1 Cs #2 Cs #3 Cs #4 Cs #5
Cs #6 Cs #7 Cs #8 C1 #1 C1 #2
C1 #3 C1 #4 C1 #5 C1 #6 C1 #7
C1 #8 C2v #1 C2 #1 Cs #9 Cs #10
respectively. Only C1 #9 is stable, with C1 #10 and #11 coalescing into C1 #6 and
#10, respectively. The bond breaking in Cs #10 suggests that we try three additional
Cs structures (#11-#13) in which the in plane OH hydroxyls have rotated 180
degrees. To our surprise, Cs #11 and #13 resulted in inversion about the bridging
oxygen. These structures, in turn, suggested trying two other Cs structures (#14 and
#15), keeping the inverted oxygen but placing the hydrogen in its original position.
None of these Cs structures were stable. Desymmetrization of Cs #11–15 to C1 #12–
16 gives stable structures at some levels, or coalescence to others.
3.4.7 [B2O(OH)4]0
A dehydrated form of boric acid dimer with a bridging oxo group was considered
(Fig. 9). The high-symmetry structures D2h #1 and #2 has B2u, Au, and B3u
A Crystallographic Review of Alkali Borate Salts … 133
C2 #2 C2 #3 Cs/1 #6
3.4.8 [B2O2(OH)4]2–
Structures containing two oxo bridges between the two boron atoms are now
considered. The first such structure, [B2O2(OH)4]2–, can be considered to be the
dimer of the hypothetical oxodihydroxoborate ion, [BO(OH)2]– (Fig. 10). First, the
high symmetry D2h #1 and #2 structures were considered. These had imaginary Au,
B3g, B2g, and (#2) B1u modes, leading to the potential D2 #1/2, C2h #1/3, C2h #2/4,
and C2v structures.
The D2 #1 structure is stable at all levels, whereas D2 #2 is only stable at
B3LYP/6-31G* and MP2/6-31G*. It has imaginary B2 and B1 modes at the other
levels, giving stable C2 #3 and #4 structures. The C2v structure, derived from D2h
#2, ascends in symmetry to the D2h #1 structure. The C2h #1 structure is stable at all
levels except B3LYP/6-311+G* and the HF levels, for which it has an Au mode to
desymmetrize to the putative C2 #1 structure, which ascends in symmetry to D2 #1.
The C2h #2 structure, which exists at all levels except HF/6-31+G* and HF/6-311
+G*, has an Au and Bg mode at all levels, suggesting desymmetrization to C2 #2
and Ci #1, respectively. These ascend in symmetry to D2 #1 and C2h #1, respec-
tively. The C2h #3 structure, which exists for calculations with diffuse basis sets, has
imaginary Bg and Bu modes, desymmetrizing to Ci #2 and Cs #1, respectively,
which are stable at all levels, with the exception of Cs #1 at HF/6-31+G*, which
ascends in symmetry via C1 #1 to C2 #3. The C2h #4 structure coalesced with either
the C2h #2 or D2h #1 structure.
C2h #1 C2h #2 Cs #1 Ci #2 D2 #2
C2h C2v C2 #3 C2 #4
C2v #1 Cs #2 C1 #4 C2 #2 C2v #2
C2 #1 Cs #1 C1 #1 C1 #5 Cs #4
3.4.9 [B2O2(OH)4]−
The diborate ion, [B2O2(OH)3]−, is shown in Fig. 10. Initially, two C2v structures
were tried. Both had imaginary B1, B2, and A2 modes. These lead to Cs #1/3, Cs #2/
4, and C2 #1/2, respectively. The structure Cs #3 coalesced into Cs #1. None of
these structures were stable. The structure Cs #1 desymmetrized to C1 #1. The
structure Cs #2, desymmetrized via C1 #2 to coalesce with C1 #1. The structure C2
#1 also desymmetrized via C1 #3 to coalesce with C1 #1. The structure C2 #2
desymmetrized to C1 #4. The structure Cs #4 desymmetrized via C1 #5 to coalesce
to C1 #4 (except at MP2/6-311+G*).
3.4.10 [B2O2(OH)4]0
The diboric acid, [B2O2(OH)2]0, which is an isomer of metaboric acid, is also given
in Fig. 10. The two forms examined, C2h and C2v, were both minima, with C2h
slightly lower in energy.
1.4902 1.488 1.485 1.515 1.509 1.501 1.499 1.505 1.503 1.528
1.5075 1.515 1.514 1.518 1.520 1.527 1.534 1.529 1.530 1.533
B5O6(OH)−
4 D2d #1 [34] 1.356 1.332 1.333 1.331 1.347 1.348 1.346 1.352 1.355 1.350
1.367 1.371 1.371 1.370 1.386 1.386 1.384 1.391 1.393 1.387
1.383 1.375 1.374 1.372 1.389 1.390 1.387 1.393 1.394 1.388
1.468 1.464 1.464 1.464 1.476 1.477 1.478 1.478 1.479 1.476
137
138 C. C. Pye
4 Conclusions
References
1. Pye CC (2018) An ab initio study of boric acid, borate, and their interconversion. Prog Theor
Chem Phys 31:143–177
2. Konig H, Hoppe R (1977) Zur Kenntnis von Na3BO3. Z Anorg Allg Chem 434:225–232 (in
German)
3. Menchetti S, Sabelli C (1982) Structure of hydrated sodium borate Na2[BO2(OH)]. Acta
Crystallogr B 38:1282–1284
4. Block S, Perloff A (1963) The direct determination of the crystal structure of NaB
(OH)4.2H2O. Acta Crystallogr 16:1233–1238
5. Csetenyi LJ, Glasser FP, Howie RA (1993) Structure of sodium tetrahydroxyborate. Acta
Crystallogr C 49:1039–1041
6. Andrieux J, Goutaudier C, Laversenne L, Jeanneau E, Miele P (2010) Synthesis,
characterization, and crystal structure of a new trisodium triborate, Na3[B3O4(OH)4]. Inorg
Chem 49:4830–4835
7. Corazza E, Menchetti S, Sabelli C (1975) The crystal structure of Na3[B3O5(OH)2]. Acta
Crystallogr B 31:1993–1997
8. Marezio M, Plettinger HA, Zachariasen WH (1963) The bond lengths in the sodium
metaborate structure. Acta Crystallogr 16:594–595
9. Menchetti S, Sabelli C (1979) A new borate polyanion in the structure of Na8[B12O20(OH)4].
Acta Crystallogr B 35:2488–2493
10. Menchetti S, Sabelli C (1977) The crystal structure of synthetic sodium pentaborate
monohydrate. Acta Crystallogr B 33:3730–3733
11. Menchetti S, Sabelli C, Trosti-Ferroni R (1982) The structure of sodium borate Na3[B5O9].
H2O. Acta Crystallogr B 38:2987–2991
12. Font Tullot JM (1947) El borax como estructura cristalina en cadenas. Estud Geol 7:13–20
(in Spanish)
13. Norimoto N (1956) The crystal structure of borax. Mineral J 2:1–18
14. Levy HA, Lisensky GC (1978) Crystal structures of sodium sulfate decahydrate (Glauber’s
Salt) and sodium tetraborate decahydrate (Borax). Redetermination by neutron diffraction.
Acta Crystallogr B 34:3502–3510
15. Gainsford GJ, Kemmitt T, Higham C (2008) Redetermination of the borax structure from
laboratory X-ray data at 145 K. Acta Crystallogr E 64:i24–i25
A Crystallographic Review of Alkali Borate Salts … 139
42. Zhang H-X, Zhang J, Zheng S-T, Yang G-Y (2005) Two new potassium borates,
K4B10O15(OH)4 with stepped chain and KB5O7(OH)2•H2O with double helical chain. Cryst
Growth Des 5:157–161
43. Salentine CG (1987) Synthesis, characterization, and crystal structure of a new potassium
borate, KB3O5•3H2O. Inorg Chem 26:128–132
44. Wang G-M, Sun Y-Q, Zheng S-T, Yang G-Y (2006) Synthesis and crystal structure of a
novel potassium borate with an unprecedented [B12O16(OH)8]4− anion. Z Anorg Allg Chem
632:1586–1590
45. Li H-J, Liu Z-H, Sun L-M (2007) Synthesis, crystal structure, vibrational spectroscopy and
thermal behavior of the first alkali metal hydrated hexaborate: K2[B6O9(OH)2]. Chin J Chem
25:1131–1134
46. Bubnova RS, Fundamenskii VS, Filatov SK, Polyakova IG (2004 ) Crystal structure and
thermal behavior of KB3O5. Doklady Phys Chem 398:249–253
47. Krogh-Moe J (1974) The crystal structure of pentapotassium enneakaidekaborate,
5K2O.19B2O3. Acta Crystallogr B 30:1827–1832
48. Zachariasen WH (1937) The crystal structure of potassium acid dihydronium pentaborate
KH2(H3O)2B5O10, (Potassium Pentaborate Tetrahydrate). Z Kristallogr 98:266–274
49. Cook WR Jr, Jaffe H (1957) The crystallographic, elastic, and piezoelectric properties of
ammonium pentaborate and potassium pentaborate. Acta Crystallogr 10:705–707
50. Zachariasen WH, Plettinger HA (1963) Refinement of the structure of potassium pentaborate
tetrahydrate. Acta Crystallogr 16:376–379
51. Gao Y-H (2011) Potassium decaborate monohydrate. Acta Crystallogr E 67:i57
52. Wu Q (2011) Potassium pentaborate. Acta Crystallogr E 67:i67
53. Krogh-Moe J (1959) On the structural relationship of vitreous potassium pentaborate to the
crystalline modifications. Arkiv Kemi 14:567–572
54. Krogh-Moe J (1972) The crystal structure of the high-temperature modification of potassium
pentaborate. Acta Crystallogr B 28:168–172
55. Krogh-Moe J (1959) The crystal structures of potassium pentaborate, K2O.5B2O3, and the
isomorphous rubidium compound. Arkiv Kemi 14:439–449
56. Krogh-Moe J (1965) Least-squares refinement of the crystal structure of potassium
pentaborate. Acta Crystallogr 18:1088–1089
57. He M, Okudera H, Simon A, Kohler J, Jin S, Chen X (2013) Structure of Li4B2O5:
high-temperature monoclinic and low-temperature orthorhombic forms. J Solid State Chem
197:466–470
58. Rousse G, Baptiste B, Lelong G (2014) Crystal structures of Li6B4O9 and Li3B11O18 and
application of the dimensional reduction formalism to lithium borates. Inorg Chem 53:6034–
6041
59. Touboul M, Betourne E, Nowogrocki G (1995) Crystal structure and dehydration process of
Li(H2O)4B(OH)4•2H2O. J Solid State Chem 115:549–553
60. Zachariasen WH (1964) The crystal structure of lithium metaborate. Acta Crystallogr
17:749–751
61. Hohne E (1964) Die Kristallstruktur des LiB(OH)4. Z Chem 4:431–432 (in German)
62. Fronczek FR, Aubry DA, Stanley GG (2001) Refinement of lithium tetrahydroxoborate with
low-temperature CCD data. Acta Crystallogr E 57:i62–i63
63. Marezio M, Remeika JP (1966) Polymorphism of LiMO2 compounds and high-pressure
single-crystal synthesis of LiBO2. J Chem Phys 44:3348–3353
64. Louer D, Louer M, Touboul M (1992) Crystal structure determination of lithium diborate
hydrate, LiB2O3(OH).H2O, from X-ray powder diffraction data collected with a curved
position-sensitive detector. J Appl Crystallogr 25:617–623
65. Krogh-Moe J (1962) The crystal structure of lithium diborate, Li2O.2B2O3. Acta Crystallogr
15:190–193
66. Krogh-Moe J (1968) Refinement of the crystal structure of lithium diborate, Li2O.2B2O3.
Acta Crystallogr B 24:179–181
A Crystallographic Review of Alkali Borate Salts … 141
67. Natarajan M, Faggiani R, Brown ID (1979) Dilithium tetraborate, Li2B4O7. Cryst Struct
Commun 8:367–370
68. Radaev SF, Muradyan LA, Malakhova LF, Burak YV, Simonov VI (1989) Atomic Structure
and electron density of lithium tetraborate Li2B4O7. Sov Phys Crystallogr 34:842–845
69. Jiang A, Lei S, Huang Q, Chen T, Ke D (1990) Structure of lithium heptaborate, Li3B7O12.
Acta Crystallogr C 46:1999–2001
70. Radaev SF, Genkina EA, Lomonov VA, Maksimov BA, Pisarevskii YV, Chelokov MN,
Simonov VI (1991) Distribution of deformation electron density in lithium triborate LiB3O5.
Sov Phys Crystallogr 36:803–807
71. Radaev SF, Maximov BA, Simonov VI, Andreev BV, D’Yakov VA (1992) Deformation
density in lithium triborate, LiB3O5. Acta Crystallogr B 48:154–160
72. Le Henaff C, Hansen NK, Protas J, Marnier G (1997) Electron density distribution in LiB3O5
at 293 K. Acta Crystallogr B 53:870–879
73. Sennova N, Albert B, Bubnova R, Krzhizhanovskaya M, Filatov S (2014) Anhydrous
lithium borate, Li3B11O18, crystal structure, phase transition and thermal expansion.
Z Kristallogr 229(7):497–504
74. Cardenas A, Solans J, Byrappa K, Shekar KVK (1993) Structure of lithium catena-Poly
[3,4-dihydroxopentaborate-1:5-mu-oxo]. Acta Crystallogr C 49:645–647
75. Zviedre I, Ievins A (1974) Crystalline structure of rubidium monoborate RbBO2•4/3 H2O.
Latvijas PSR Zinatnu Akademijas Vestis Kimijas Seria 4:395–400 (in Russian)
76. Schmid S, Schnick W (2004) Rubidium metaborate, Rb3B3O6. Acta Crystallogr C 60:i69–i70
77. Touboul M, Penin N, Nowogrocki G (2000) Crystal structure and thermal behavior of
Rb2[B4O5(OH)4]•3.6H2O. J Solid State Chem 149:197–202
78. Krzhizhanovskaya MG, Bubnova RS, Bannova II, Filatov SK (1997) Crystal structure of
Rb2B4O7. Crystallogr Rep 42:226–231
79. Bubnova RS, Krivovichev SV, Shakhverdova IP, Filatov SK, Burns PC,
Krzhizhanovskaya MG, Polyakova IG (2002) Synthesis, crystal structure and thermal
behavior of Rb3B7O12, a new compound. Solid State Sci 4:985–992
80. Krzhizhanovskaya MG, Kabalov YK, Bubnova RS, Sokolova EV, Filatov SK (2000)
Crystal structure of the low-temperature modification of α-RbB3O5. Crystallogr Rep 45:572–
577
81. Krzhizhanovskaya MG, Bubnova RS, Fundamentskii VS, Bannova II, Polyakova IG,
Filatov SK (1998) Crystal structure and thermal expansion of high-temperature β-RbB3O5
modification. Crystallogr Rep 43:21–25
82. Liu Z-H, Li L-Q, Zhang W-J (2006) Two new borates containing the first examples of large
isolated polyborate anions: chain [B7O9(OH)5]2− and ring [B14O20(OH)6]4−. Inorg Chem
45:1430–1432
83. Behm H (1984) Rubidium pentaborate tetrahydrate, Rb[B5O6(OH)4].2H2O. Acta Crystal-
logr C 40:217–220
84. Penin N, Seguin L, Touboul M, Nowogrocki G (2001) Crystal structures of three MB5O8
(M = Cs, Rb) borates (α-CsB5O8, γ-CsB5O8, and β-RbB5O8). J Solid State Chem 161:205–213
85. Schlager M, Hoppe R (1994) Darstellung und Kristallstruktur von CsBO2. Z Anorg Allg
Chem 620:1867–1871 (in German)
86. Touboul M, Penin N, Nowogrocki G (1999) Crystal structure and thermal behavior of
Cs2[B4O5(OH)4] 3H2O. J Solid State Chem 143:260–265
87. Nowogrocki G, Penin N, Touboul M (2003) Crystal structure of Cs3B7O12 containing a new
large polyanion with 63 boron atoms. Solid State Sci 5:795–803
88. Krogh-Moe J (1960) The crystal structure of cesium triborate, Cs2O.3B2O3. Acta Crystallogr
13:889–892
89. Krogh-Moe J (1974) Refinement of the crystal structure of caesium triborate, Cs2O.3B2O3.
Acta Crystallogr B 30:1178–1180
90. Penin N, Seguin L, Touboul M, Nowogrocki G (2002) A new cesium borate Cs3B13O21.
Solid State Sci 4:67–76
142 C. C. Pye
Cory C. Pye
Abstract The chemistry of boric acid and monomeric borates is reviewed. Fol-
lowing a discussion of the crystal structures and nuclear magnetic resonance
studies, ab initio results are presented of molecular ortho- and metaboric acid,
(tetrahydroxo)borate, and the hydrates of orthoboric acid and borate. The structures
and vibrational frequencies are compared with experiment. Attempts to study their
interconversion lead us to a discussion of oxodihydroxoborate (the conjugate base
of boric acid), and of the hydroxide-boric acid complex. It is hypothesized that the
conversion of boric acid into borate proceeds via the oxodihydroxoborate inter-
mediate. Finally, the calculated structures of hydroxodioxo- and trioxoborate are
compared with experiment.
1 Introduction
C. C. Pye (✉)
Department of Chemistry, Saint Mary’s University, Halifax, NS B3H 3C3, Canada
e-mail: [email protected]
(continued)
Table 1 (continued)
146
Stoichiometry Name Synth Space Unit cell X-ray B-O (Å) ranges Comments
BmHnO(3m+n)/2 T(K) group (Å or deg.) T(K)
m n Tetrahedral Trigonal References
423 P21 ̸ c 6.758,8.844, 183 1.3284– 1.3485– [27]
7.075 93.5 1.5612 1.3812
Metaboric acid P43n 8.886 1.436–1.505 n/a [28]
(cubic γ)
453 P43n 8.8811 183 1.4428– n/a [27]
1.5094
C. C. Pye
Table 2 Crystal structures of monomeric sodium and lithium borates
Stoichiometry Synth Space Unit cell X-ray B-O (Å) ranges Comments
MlBmHnO(l+3m+n)/2 T(K) group (Å or deg.) T(K)
M l m n Tetrahedral Trigonal Ref.
Na 3 1 0 748 P21 ̸ c 5.687,7.530,9.993 n/a 1.377–1.409 [BO3]3− ions [29]
127.15
Na 2 1 1 523 Pnma 8.627,3.512,9.863 n/a 1.351–1.439 [BO2(OH)]2− ions [30]
An Ab Initio Study of Boric Acid, Borate …
Early work on the nuclear magnetic resonance (NMR) spectra of boron com-
pounds (11B-12.83 MHz) was done and showed a range of chemical shifts and
some 1-bond 11B-1H and 11B-2H coupling constants, using BF3 ⋅ Et2O as a ref-
erence [37]. Of interest in this study are the 11B chemical shifts of NaBO2(aq)
(−1.3 ± 0.5 ppm, due to [B(OH)4]−), NaB5O8(aq) (−1.3 and −14.4 ± 1.0 ppm),
K2B4O7 (−7.5 ± 1.0 ppm), Na2B4O7 (−8.0 ± 0.5 ppm), (NH4)2B4O7 (−10.3 ±
0.5 ppm), KB5O8(aq) (−13.0 ± 0.5 ppm), and B(OH)3(aq) (−18.8 ± 1.0 ppm).
These were interpreted as being due to a dynamic equilibrium between B(OH)4−
and B(OH)3. A later study by Momii and Nachtrieb (sat. B(OH)3 (aq) reference,
11
B-14 MHz) reexamined these results and gave 0.090–0.900 M NaBO2(aq) at
17.4 ± 0.5 ppm and 0.090–0.900 M KBO2 at 15.5 ± 0.5 ppm [38]. These were
interpreted as due to [B(OH)4]−. For sodium pentaborate solutions, the peak at
15.0 ppm was assigned to [B5O6(OH)4]−, and the peak at 1.1 ppm assigned to a
rapid equilibrium between B(OH)3, B(OH)4− and B3O3(OH)4−. For the tetrabo-
rates, a single peak is observed whose chemical shift increases with concentration
from 8 to 11 ppm, and this was assigned to a rapid equilibrium between B(OH)3,
B(OH)4− and at least two other ions. How and coworkers showed that the chemical
shift of a 50 g/L solution at 33 °C varied from −2 to −20 ppm between pH 12–2
respectively (11B, 12.83 MHz, BF3 ⋅ Me2O ref.) [39]. Smith and Wiersema
(11B-80 MHz) noted that one NMR peak in all borate solutions was linearly related
to the sodium to boron ratio and could this be interpreted as the peak of rapidly
exchanging B(OH)3 and [B(OH)4]− [40]. For tetraborate solutions, three peaks
could be observed, with the 5.0 ppm peak assigned to [B3O3(OH)4]−. Pentaborate
solutions also showed three peaks, with the 5 ppm peak assigned to [B3O3(OH)4]−,
and the peak at 18 ppm assigned to [B5O6(OH)4]−. Covington and Newman
examined the 11B spectra (28.87 MHz, rel. to infinite dilution [B(OH)4]−) of
sodium and potassium borate in water and in ∼0.1 mol/L added [OH−] in an effort
to determine the pKb of borate [41]. Henderson et al., in their study of the com-
plexation of borate with diols, showed the 11B NMR (12.83 MHz) of borax, boric
acid, and sodium metaborate from pH 2–12, along with the line width at half height
[42]. Janda and Heller examined the 11B spectra (60 MHz) of sodium, potassium,
and ammonium polyborates as a function of concentration and pH (0.5–13.8) and
either one or two lines were observed [43]. Epperlein et al. examined the 10B
spectra (1.807T, 8.267 MHz) of some boron species and found B(OH)3 at 0 ppm
(reference), B5O6(OH)4− at 17 ppm, B4O5(OH)42− between 70–85 ppm, and B
(OH)4− at around 140 ppm [44]. Salentine confirmed earlier results (11B, 127 and
160 MHz, external reference BF3 ⋅ Et2O) on the pentaborate (18, 13, 1 ppm) and
tetraborate (12, 8, 1 ppm) [45]. It was proposed that the resonance at 13 ppm, due
to triborate ion, and at 1 ppm, due to pentaborate ion, are due to the tetrahedral
boron atoms, and the trigonal boron atoms are not observed because of quadrupolar
relaxation.
We have reviewed the crystallography of boric acid and monomeric borates, and
the boron NMR of boric-acid/borate containing solutions. Our remaining goals are
to compare the ab initio energy, structure, and vibrational spectra to experiment
(where known) of orthoboric acid, metaboric acid, and tetrahydroxoborate, and to
An Ab Initio Study of Boric Acid, Borate … 149
determine the effect of hydration where appropriate. We also examine the mecha-
nism of interconversion of orthoboric acid and tetrahydroxoborate, which leads to a
discussion of the crystallography unobserved oxodihydroxoborate anion. We then
discuss the related crystallographically observed trioxoborate and dioxohydroxob-
orate anions.
2 Methods
Calculations were performed using Gaussian 03 [46]. The MP2 calculations utilize
the frozen core approximation, which is valid in nearly all cases except where
excessive core/valence mixing occurs (denoted c/v). The geometries were opti-
mized using a stepping stone approach, in which the geometries at the levels HF/
6-31G*, HF/6-31+G*, HF/6-311+G*, B3LYP/6-31G*, B3LYP/6-31+G*, B3LYP/
6-311+G*, MP2/6-31G*, MP2/6-31+G*and MP2/6-311+G* were sequentially
optimized, with the geometry and molecular orbital reused for the subsequent level.
Default optimization specifications were normally used. After each level, where
possible, a frequency calculation was performed at the same level and the resulting
Hessian was used in the following optimization. Z-matrix coordinates constrained
to the appropriate symmetry were used as required to speed up the optimizations.
Because frequency calculations are done at each level, any problems with the
Z-matrix coordinates would manifest themselves by giving imaginary frequencies
corresponding to modes orthogonal to the spanned Z-matrix space. The Hessian
was evaluated at the first geometry (Opt = CalcFC) for the first level in a series in
order to aid geometry convergence. To facilitate comparison with results from
gas-phase, solution, and solid phase measurements, no solvent corrections were
applied except via the supermolecule approach (explicit water molecules).
Six forms of (ortho)boric acid were investigated (Fig. 1). Two of these (C3h and Cs #2)
were minima, with the C3h structure being lower in energy by 22–26 kJ/mol
(Table 3). The C3v structure, a third-order saddle point, was much higher in energy
(139–152 kJ/mol). The other three structures were transition states linking the minima.
Both the Cs #1 and Cs #3 linked the Cs #2 structure to itself, whereas the C1 structure
linked the C3h and Cs #2 structures. The C1 structure was 38–43 kJ/mol higher than
the C3h structure, whereas the Cs #1 and Cs #3 structures were 34–38 and
20–22 kJ/mol, respectively, higher than the Cs #2.
150 C. C. Pye
C3h C3v Cs #1
Cs #2 Cs #3 C1 #1
Fig. 1 Structure of boric acid, [B(OH)3]. A bold symmetry label indicates a minimum energy
structure
Of these structures, the C3h structure is observed in the crystal structure of both
polymorphs of boric acid (Table 1). Our calculated B-O bond lengths range from
1.3553 to 1.3780 Å, with HF < B3LYP < MP2 (Table 4). These are in good
agreement with the experimentally determined (X-ray) average bond length of
1.36 Å and with previous literature values (Tables 1 and 4).
The vibrational spectra (unscaled) of the lowest-energy form of boric acid, as
well as some experimental vibrational frequencies from the literature, is given in
Table 5. Undistorted boric acid, of C3h symmetry, has 15 modes of internal
vibration and spans the vibrational representation
Table 4 Geometrical Parameters of the C3h form of B(OH)3. n/r = not reported
Level B-O (Å) O-H (Å) B-O-H angle (deg.)
HF/6-31G* 1.3581 0.9466 112.56
HF/6-31+G* 1.3584 0.9471 113.69
HF/6-311+G* 1.3553 0.9401 113.99
B3LYP/6-31G* 1.3721 0.9675 110.90
B3LYP/6-31+G* 1.3729 0.9683 112.46
B3LYP/6-311+G* 1.3691 0.9630 113.01
MP2/6-31G* 1.3762 0.9700 110.38
MP2/6-31+G* 1.3780 0.9722 111.77
MP2/6-311+G* 1.3709 0.9614 112.25
Literature
HF/STO-3G [2, 47] 1.389 n/r 114 (fixed)
HF/STO-3G [48] 1.364 0.98 110
HF/4-31G [48] 1.364 0.95 121
HF/3-21G* [49] 1.377 0.962 n/r
HF/6-31G [49] 1.370 0.947 n/r
HF/6-31G* [49] 1.358 0.947 n/r
HF/6-31G* [2, 49] 1.358 0.947 112.6
MP2/6-31G** [50] 1.357 0.942 113.0
B3LYP/6-311++G** [51] 1.380 0.971 112.6
B3LYP/6-311++G** [52] 1.370 0.962 112.8
MP2/6-311++G** [52] 1.373 0.961 110.7
B3LYP/aug-cc-pVQZ [53] 1.369 0.960 113.1
MP2/aug-cc-pVTZ [53] 1.374 0.962 111.5
MP2/aug-cc-pVQZ [53] 1.370 0.959 111.8
QCISD/6-311++G** [53] 1.371 0.959 111.1
B3LYP/aug-cc-pVDZ [54] 1.376 n/r n/r
agreement with the IR spectra measured in an argon matrix, and with most modes
of solid and aqueous boric acid. However, the in-plane BO3 deformation, HOB
deformation, and OH stretching frequencies differ, which would be expected,
because in solid boric acid, the molecules are held together by a network of
hydrogen bonds. It might be expected that aqueous solutions may exhibit similar
behavior to the solid. The vibrational spectra of the higher-energy Cs #2 confor-
mation is also given in Table 6. The E modes correlate with 2A modes of the same
reflection symmetry. The main differences in the vibrational frequencies of the Cs
#2 conformer are that the BOH torsion is much lower, the in plane BO3 deformation
is somewhat higher in frequency, the out of plane BO3 deformation is slightly
lower, the symmetric BO stretch is slightly higher, one component of the HOB
deformation and asymmetric B-O stretch is somewhat lower, and the OH fre-
quencies slightly higher.
Table 5 Theoretical and literature experimental vibrational frequencies (cm−1) of the C3h form of boric acid, B(OH)3
152
BO3 def BOH BOH tors BO3 oop def BO str HOB HOB BO str OH str OH str
tors def def
E′ A″ E″ A″ A′ E′ A′ E′ E′ A′
Theory, C3h
HF/6-31G* 458 476 567 736 927 1110 1116 1561 4135 4138
HF/6-31+G* 452 484 572 741 924 1086 1092 1540 4129 4132
HF/6-311+G* 454 491 577 747 919 1098 1112 1533 4188 4191
B3LYP/6-31G* 428 428 535 665 877 1038 1045 1478 3800 3800
B3LYP/6-31+G* 420 443 543 669 871 1010 1015 1448 3796 3796
B3LYP/6-311+G* 420 446 544 676 869 1018 1029 1441 3829 3830
MP2/6-31G* 425 445 551 675 880 1052 1056 1494 3823 3823
MP2/6-31+G* 419 450 551 677 872 1021 1024 1458 3793 3794
MP2/6-311+G* 426 447 548 680 873 1031 1038 1459 3872 3873
Theory (Literature)
MP2/6-31G** [50] 428 445 547 682 886 1044 1045 1509 3949 3948
B3LYP/6-311++G** 427 473 669 1015 1442 3870
[51]
Experiments
Raman, sat.aq. [55] 875
Raman, powder [56] 880 3172 3256
Raman, solution [57] 506 875 1130? 1060 1420
Raman, powder 503 1155 1065
Raman, sol. 70 °C [58] 882 3195 3290
Raman, powder 515 872 986
Raman, powder [59] 883 1169 1032 1391 3170 3248
(continued)
C. C. Pye
Table 5 (continued)
BO3 def BOH BOH tors BO3 oop def BO str HOB HOB BO str OH str OH str
tors def def
E′ A″ E″ A″ A′ E′ A′ E′ E′ A′
IR, mull [60] 885 1195 1450 3270
IR, mull and xtal. [61] 540 648 882 1197 1450 3200
IR, KBr pellets [62] 544 639 882? 1183 1428 3150
IR, 10B(OH)3 545 668 882? 1195 1490
IR, B(OD)3 540 654 1428 2380
Raman, sol 500 880 1430 3200
R, pH = 6.1–10 [63] 875
R, pH = 6.38–10.60 [64] 872
An Ab Initio Study of Boric Acid, Borate …
Table 6 Theoretical vibrational frequencies (cm−1) of the Cs form of boric acid, B(OH)3
BO3 BOH BOH BO3 BO HOB HOB BO OH OH
def tors tors oop str def def str str str
def
A′ A″ A″ A″ A′ A′ A′ A′ A′ A′
HF/6-31G* 471 306 508 729 930 1006 1131 1517 4141 4168
488 549 1116 1576 4156
HF/6-31+G* 468 324 514 733 928 979 1099 1495 4136 4166
483 558 1093 1554 4154
HF/6-311+G* 471 313 506 735 923 992 1111 1488 4195 4221
486 560 1107 1543 4211
B3LYP/6-31G* 441 260 476 661 877 947 1058 1430 3808 3830
458 527 1042 1494 3814
B3LYP/6-31+G* 436 295 482 665 873 912 1022 1400 3806 3833
450 539 1012 1463 3816
B3LYP/6-311+G* 438 274 466 668 871 926 1031 1392 3838 3862
452 536 1019 1454 3847
MP2/6-31G* 439 274 491 670 879 963 1074 1446 3825 3853
457 542 1057 1510 3842
MP2/6-31+G* 435 289 487 672 872 928 1036 1410 3797 3830
451 546 1024 1473 3819
MP2/6-311+G* 443 262 467 673 874 942 1047 1411 3875 3904
459 540 1037 1472 3894
Four Cs forms of hydrated boric acid with a water molecule directly bound to the
boron atom were attempted. Alternatively, this may be viewed as protonated
monoborate anion. In all cases, the water molecule dissociated. In the first two
cases, a solvated boric acid was obtained, neither of which was an energy minimum
(based on Cs #1 and Cs #3 naked boric acid, see Fig. 3). In the final two cases,
An Ab Initio Study of Boric Acid, Borate … 155
C3h C3v Cs #1
Cs #2 C s #3 C1 #1
z-matrix errors occurred as the water molecule rotated. It can be said therefore that
direct protonation of B(OH)4− results in water elimination.
Four forms of hydrated boric acid with a water molecule hydrogen-bonded to the
boric acid were attempted (Fig. 3). To the naked C3h structure, a water molecule
may only hydrogen bond in a donor-acceptor (DA) fashion (C1 #2), whereas to the
naked Cs #2 structure, a water molecule may hydrogen bond in either a double
donor (DD, Cs #1 or C1 #3), donor acceptor (DA, C1 #1), or double acceptor (AA,
Cs #2). All were energy minima, with the exception of Cs #1 at the B3LYP/6-31G*
and all MP2 levels. These desymmetrized to C1 #3, where the water is removed
from the BO3 plane by varying amounts (significantly at the B3LYP and
MP2/6-31G* levels).
For the dihydrate, a Cs and five C1 forms were tried. The C1 #4 and C1 #5 forms
are derived from the naked C3h structure and differ in the orientation of the free
hydrogen of the water molecules, either up,up (uu) or up,down (ud). The other
forms are derived from the naked Cs #2 structure: namely, the Cs/1 #1 (DD,AA), the
C1 #2 (DD,DA), and the C1 #3 (DA,AA). The Cs form was stable at the HF,
156 C. C. Pye
Cs #1 Cs #2 Cs #2a Cs #1/ C1 #3
Cs #2 C1 #1 C1 #2 Cs/1 #1 C1 #1a
C1 #2 C1 #2a C1 #2b C1 #3 C1 #4
C1 #5 C3 #1 C1 #1 C1 #1a C1 #2
Our approach to hydration of boric acid is to solvate the boric acid molecule as
completely as possible with a small number of water molecules in order to effi-
ciently model the vibrational spectra. Of course, water molecules might prefer to
hydrogen-bond to other water molecules instead of to boric acid. This work is
therefore somewhat complementary to that of Tachikawa [52], who studied similar
clusters with up to five water molecules and found several in which water molecules
were hydrogen bonded to each other.
Five forms of monoborate were investigated (Fig. 7). Two of these (D2d #2 or S4 #3,
and S4 #1) were minima, with the S4 #1 structure being lower in energy by 7–10 kJ/mol
at the Hartree-Fock levels (Table 8). The D2d #2 is only a minimum at the Hartree-Fock
levels. We confirm the presence of a second shallow minimum (S4 #3) at the correlated
levels, as first found by Stefani et al. [53] The S4 #2 structure is a transition state that
connects the D2d #2/S4 #3 and S4 #1 structures. It is 2–3 kJ/mol higher in energy than
D2d #2. A D2 structure, derived from D2d #1, ascended in symmetry to give the D2d #2
structure at all levels.
An Ab Initio Study of Boric Acid, Borate … 159
The BOH deformations appear from 1000 to 1300 cm−1, the B-O stretching
motions from 720 to 1100 cm−1, the BO4 deformations from 230 to 570 cm−1, and
the BOH torsions from 180 to 440 cm−1.
Table 10 Theoretical and literature experimental vibrational frequencies (cm−1) of the S4 #1 form of B(OH)4−
BOH BO4 BOH BO4 BOH BO4 BO4 BO BO BO BOH BOH BOH
tors def tors def tors def def str str str def def def
A B E A B E B A E B A B E
Theory
HF/6-31G* 202 273 317 420 431 514 568 775 940 1086 1110 1169 1323
HF/6-31+G* 204 260 307 420 429 511 561 774 924 1056 1086 1141 1285
HF/6-311+G* 212 272 313 429 438 512 568 770 920 1049 1098 1151 1285
B3LYP/6-31G* 190 264 331 345 416 476 521 728 866 1018 1026 1070 1252
B3LYP/6-31+G* 181 235 307 366 407 471 513 724 851 973 1000 1048 1193
B3LYP/6-311+G* 189 245 311 380 408 473 518 721 846 964 1015 1060 1195
MP2/6-31G* 203 272 343 368 423 482 535 732 892 1041 1042 1094 1272
An Ab Initio Study of Boric Acid, Borate …
MP2/6-31+G* 203 249 326 372 414 473 520 725 869 990 1012 1060 1208
MP2/6-311+G* 212 265 323 406 412 477 534 729 871 993 1016 1067 1206
Theory (Literature)
HF/6-31G* [70] 202 273 316 420 431 514 569 775 940 1086 1110 1169 1323
Experiments
NaBO2, 5N [71] 749
NaBO2 (aq) [57] 400– 747 950 1132 1276
500
NaBO2 ⋅ 4H2O(s) 741
H3BO3+NaOH 744–7 949 1150– 1219–
60 20
KB(OH)4(aq) (Td) [72] 379 379 533 533 754 947 947
NaB(OH)4(aq) (Td) [73] 374 374 516 516 744 940 940
sat NaB(OH)4(aq) pH = 5–10 [63] 745
(continued)
161
Table 10 (continued)
162
BOH BO4 BOH BO4 BOH BO4 BO4 BO BO BO BOH BOH BOH
tors def tors def tors def def str str str def def def
A B E A B E B A E B A B E
NaB(OH)4(aq), 1.5M, pH = 6–12 [64] 375 375 522 522 745 940 940
NaB(OH)4(s) [65] 405 405 520 520 738 952 952 1078
Na2B4O7(aq), 0.01–0.8 [65] 744
sat Cs2B4O7 ⋅ 5H2O or CsB5O8 ⋅ 4H2O 480,484 511 742,745
[74]
NaBO2(aq), 4–5 M [54] 740
KBO2 ⋅ nH2O(aq), n = 5, 10, 19 [75] 740
C. C. Pye
An Ab Initio Study of Boric Acid, Borate … 163
C1 C2 C1 C1 S4
The monoborate anion may be hydrated. The monohydrate has C1 symmetry; the
dihydrate, either C1 or C2; the trihydrate, C1; and the tetrahydrate, S4 (see Fig. 8).
Upon hydration, the average B-O distance slightly decreases (0.002 Å) but the
deviation from the mean can be as much as 0.025 Å (Fig. 9).
The vibrational frequencies (MP2/6-311+G*) of hydrated monoborate are
plotted in Fig. 10. The OH stretching frequencies (not shown) decrease by about
50 cm−1 upon going from the anhydrate to the tetrahydrate. The restricted trans-
lations of the water molecules appear below 250 cm−1, whereas the restricted
rotations appear at approximately 300–440 cm−1 (rock), 580–730 cm−1 (twist), and
770–820 cm−1 (wag). The BOH deformation frequencies are in the range from
1000–1220 cm−1 and tend to increase with hydration as the mode becomes stiffer
upon forming a hydrogen bond to water. The B-O totally symmetric stretch only
increases by a few wavenumbers, whereas the asymmetric stretch increases much
more. The deformation modes tend to increase a bit, but the BOH torsional modes
increase a lot because of the restrictions imposed by hydrogen bonding. There is
significant mixing both between these two types of modes and with the water
rocking modes.
Attempts to hydrate boric acid by attaching water directly to the boron (see above,
equivalent to direct protonation of monoborate) resulted in the dissociation of the
water molecule. It was thought that the more realistic hydronium ion might interact
with the borate ion by forming an ion pair before losing a water molecule. A scan of
the O…H distance from 2.0 to 0.9 Å in 0.1 Å steps was carried out in an attempt to
carefully protonate the borate at HF/6-31G*; however, a different proton from the
An Ab Initio Study of Boric Acid, Borate … 165
The dissociation of one of the B-O bonds of borate to form boric acid and
hydroxide is a simple possibility for their interconversion. Scans were done at each
level, with the zero of energy set to the optimized borate structure and are shown in
Fig. 11. Typically the scans suggest that as the B-O distance increases, two of the
three remaining borate hydroxyls rotate to form stabilizing hydrogen bonds with
the departing hydroxide, which then swings into the plane of a Cs #2 boric acid as
the transition state is passed. However, for the B3LYP/6-31G* and MP2/6-31G*
scans, which proceeded farther than the others, the hydroxide abstracted a proton
from boric acid to give BO(OH)2− + H2O. This suggested the need to explore the
relative energies of hydrated BO(OH)2− and of the boric acid-hydroxide complex.
C2v #1 C2v #2 Cs C2
HF/6-311+G* 1.2831 1.4199 455 505 542 557 792 847 1077 1183 1287 1639
B3LYP/6-31G* 1.2949 1.4455 412 465 493 517 703 791 972 1106 1235 1605
B3LYP/6-31+G* 1.3041 1.4398 412 459 512 532 708 792 976 1082 1204 1532
B3LYP/6-311+G* 1.2992 1.4379 415 460 504 530 709 787 975 1093 1200 1527
MP2/6-31G* 1.3014 1.4499 415 466 508 531 712 796 989 1117 1250 1617
MP2/6-31+G* 1.3135 1.4450 412 457 511 542 710 794 982 1087 1214 1522
MP2/6-311+G* 1.3062 1.4386 420 463 485 532 708 794 989 1094 1207 1525
167
168 C. C. Pye
C2 #1/ C1 #2 Cs #1 C1 #1 C1 #2a
C1 #3 C1 #4 C1 #5 C1 #6 C2v
C2/1 #1 Cs #1 C1 #2 C1 #3 Cs #2
observe it spectroscopically. The vibrational frequencies of the most stable form are
given in Table 12. In the isotropic Raman spectra, one would predict the obser-
vation of the BO3 deformation mode at around 415 cm−1, the BO(H) stretching
mode at around 800 cm−1, the BOH deformation mode at around 1100 cm−1, and
the BO stretching mode at about 1530 cm−1. There is a fair amount of coupling
between the BO stretching and BOH deformation modes.
A monohydrate can be based on any of the three stable anhydrous forms. From
the most stable anhydrous C2v #1 form, both the C2v #1 and the more stable C1 #1
forms can be derived (Fig. 13). The C2v #1 is only stable at the HF levels, and at
B3LYP/6-31G*, converting to the Cs #1 form at the other levels. From the next
most stable anhydrous Cs form, the Cs #2, C1 #2 and C1 #3 forms can be derived.
An Ab Initio Study of Boric Acid, Borate … 169
The boric acid-boric acid complex was investigated next (Fig. 15). All attempts to
locate the complex between hydroxide and the C3h form of boric acid resulted in
deprotonation to form the oxodihydroxoborate anion-water complex. The Cs
170 C. C. Pye
Cs C1
C1 #1 C1 #2 C1 #3 C1 #1 C1 #2
complex between the Cs #2 form of boric acid acting as a double hydrogen bond
donor to hydroxide was not stable at B3LYP/6-31G* and MP2/6-31G*, reverting
instead to a oxodihydroxoborate-water complex via proton transfer. At the other
levels, an imaginary frequency gave rise to the stable C1 structure, which was 10–
25 kJ/mol less stable than the corresponding oxodihydroxoborate form. Upon
An Ab Initio Study of Boric Acid, Borate … 171
C1 C1 #1 C1 #2 C1 #3 C1 #4
C1 #5 C1 #1 C1 #1a C1 #2 C1 #2a
C1 #5 C1 #6 C1 #7 C1 #8 C1 #8a
stabilize the transition state even more to bring the results into even better agree-
ment with experiment, assuming that the postulated intermediate is correct.
their sodium salts (Table 1). The calculated B-O bond distance of BO33− (Table 13)
is slightly longer than what is observed experimentally in the sodium salt, and this
is partly due to the neglect of the sodium counterion and crystal packing. For
BO2(OH)2−, the B-O distances for the unprotonated oxygens are in reasonable
agreement with experiment, whereas the B-O(H) distance is quite long. It is
expected that hydrogen bonding present in the crystal structure would shorten the
corresponding B-O(H) distance considerably.
The calculated vibrational frequencies of BO33− are shown in Table 14. There is
a quite surprising dependence of the level of theory and basis set. The stretching
frequencies decrease when going from Hartree-Fock to the correlated levels, as
expected, as well as when going from nondiffuse to diffuse basis sets. It is more
surprising that the deformation modes are affected even more when going from
nondiffuse to diffuse basis sets at the correlated levels. This may be an artifact of
using a correlated level on a system, which, in the gas-phase, is likely unbound with
respect to electron detachment.
The calculated vibrational frequencies of BO2(OH)2− are shown in Table 15.
The dependence of the frequency on the level of theory and basis set is not as
174 C. C. Pye
pronounced as for trioxoborate. The B-O and B-OH stretching frequencies decrease
when going from Hartree-Fock to the correlated levels, as expected. When going
from nondiffuse to diffuse basis sets, the B-O stretching frequencies decrease, but
the B-OH frequency increases. The other frequencies do not show any surprising
trends.
4 Conclusions
The calculated bond lengths and vibrational frequencies of boric acid and borate
agree fairly well with that observed experimentally and with previous calculations,
where available, when the comparison is appropriate. The oxodihydroxoborate ion
is much more stable than the boric acid-hydroxide complex, when the latter exists.
The oxodihydroxoborate ion, if it can be observed, should have a strong vibrational
band at approximately 1400–1600 cm−1. A transition state that links the tetrahy-
droxoborate to the hydrated oxodihydroxoborate ion has been found. The addition
of water molecules lowers the barrier significantly, bringing the activation energy to
closer agreement with experiment (assuming an oxodihydroxoborate intermediate).
References
31. Block S, Perloff A (1963) The direct determination of the crystal structure of NaB(OH)42H2O.
Acta Crystallogr 16:1233–1238
32. Csetenyi LJ, Glasser FP, Howie RA (1993) Structure of sodium tetrahydroxyborate. Acta
Crystallogr C 49:1039–1041
33. Touboul M, Betourne E, Nowogrocki G (1995) Crystal structure and dehydration process of
Li(H2O)4B(OH)4.2H2O. J Solid State Chem 115:549–553
34. Zachariasen WH (1964) The crystal structure of lithium metaborate. Acta Crystallogr
17:749–751
35. Hohne E (1964) Die Kristallstruktur des LiB(OH)4. Z Chem 4:431–432
36. Fronczek FR, Aubry DA, Stanley GG (2001) Refinement of lithium tetrahydroxoborate with
low-temperature CCD data. Acta Crystallogr E 57:i62–i63
37. Onak TP, Landesman H, Williams RE, Shapiro I (1959) The B11 nuclear magnetic resonance
chemical shifts and spin coupling values for various compounds. J Phys Chem 63:1533–1535
38. Momii RK, Nachtrieb NH (1967) Nuclear magnetic resonance study of borate-polyborate
equilibria in aqueous solution. Inorg Chem 6:1189–1192
39. How MJ, Kennedy GR, Mooney EF (1969) The pH dependence of the boron-11
chemical-shift of borate-boric acid solutions. J Chem Soc D Chem Commun 267–268
40. Smith HDJ, Wiersema RJ (1972) Boron-11 nuclear magnetic resonance study of polyborate
ions in solution. Inorg Chem 11:1152–1154
41. Covington AK, Newman KE (1973) Base dissociation constant of the borate ion from 11B
chemical shifts. J Inorg Nucl Chem 35:3257–3262
42. Henderson WG, How MJ, Kennedy GR, Mooney EF (1973) The interconversion of aqueous
boron species and the interaction of borate with diols: a 11B N.M.R. study. Carbohydrate Res
28:1–12
43. Janda R, Heller G (1979) 11B–NMR-spektroskopische Untersuchungen an waessrigen
Polyboratloesungen. Z Naturforsch B 34:1078–1083
44. Epperlein BW, Lutz O, Schwenk A (1975) Fourier-Kernresonanzuntersuchungen an 10B und
11B in Waessriger Loesung. Z Naturforsch A 30:955–958
45. Salentine CG (1983) High-field 11B NMR of alkali borates. Aqueous polyborate equilibria.
Inorg Chem 22:3920–3924
46. Frisch MJ et al (2004) Gaussian 03, Revision D.02. Gaussian Inc., Wallingford, CT
47. Gupta A, Tossell JA (1981) A theoretical study of bond distances, X-ray spectra and electron
density distributions in borate polyhedra. Phys Chem Miner 7:159–164
48. Gupta A, Tossell JA (1983) Quantum mechanical studies of distortions and polymerization of
borate polyhedra. Am Miner 68:989–995
49. Zhang ZG, Boisen MBJ, Finger LW, Gibbs GV (1985) Molecular mimicry of the geometry
and charge density distribution of polyanions in borate minerals. Am Miner 70:1238–1247
50. Zaki K, Pouchan C (1995) Vibrational analysis of orthoboric acid H3BO3 from ab initio
second-order perturbation calculations. Chem Phys Lett 236:184–188
51. Tian SX, Xu KZ, Huang M-B, Chen XJ, Yang JL, Jia CC. Theoretical study on infrared
vibrational spectra of boric-acid in gas-phase using density functional methods. J Mol Struct
(Theochem) 459:223–227, 459
52. Tachikawa M (2004) A density functional study on hydrated clusters of orthoboric acid, B
(OH)3(H2O)n (n = 1–5). J Mol Struct (Theochem) 710:139–150
53. Stefani D, Pashalidis I, Nicolaides AV (2008) A computational study of the conformations of
the boric acid (B(OH)3), its conjugate base ((HO)2BO–) and borate anion (B(OH)4–). J Mol
Struct (Theochem) 853:33–38
54. Zhou Y, Fang C, Fang Y, Zhu F (2011) Polyborates in aqueous borate solution: a Raman and
DFT theory investigation. Spectrochim Acta A 83:82–87
55. Ananthakrishnan R (1936) The Raman spectra of some boron compounds (methyl borate,
ethyl borate, boron tri-bromide and boric acid). Proc Indian Acad Sci A 4:74–81
56. Ananthakrishnan R (1937) The Raman spectra of crystal powders. IV. Some organic and
inorganic compounds. Proc Indian Acad Sci A 5:200–221
57. Hibben JH (1938) The constitution of some boric oxide compounds. Am J Sci A 35:113–125
An Ab Initio Study of Boric Acid, Borate … 177
58. Mitra SM (1938) Raman effect in boric acid and in some boron compounds. Ind J Phys
12:9–14
59. Kahovec L (1938) Studien zum Raman-Effekt. Mitteilung LXXXV. Borsauere und Derivate.
Z Phys Chem 40:135–145
60. Miller FA, Wilkins CH (1952) Infrared spectra and characteristic frequencies of inorganic
ions. Anal Chem 24:1253–1294
61. Bethell DE, Sheppard N (1955) The infra-red spectrum and structure of boric acid. Trans
Faraday Soc 51:9–15
62. Servoss RR, Clark HM (1957) Vibrational spectra of normal and isotopically labeled boric
acid. J Chem Phys 26:1175–1178
63. Maya L (1976) Identification of polyborate and fluoropolyborate ions in solution by Raman
spectroscopy. Inorg Chem 15:2179–2184
64. Maeda M, Hirao T, Kotaka M, Kakihana H (1979) Raman spectra of polyborate ions in
aqueous solution. J Inorg Nucl Chem 41:1217–1220
65. Janda R, Heller G (1979) Ramanspektroskopische Untersuchungen an festen und in Wasser
geloesten Polyboraten. Z Naturforsch B 34:585–590
66. Ogden JS, Young NA (1988) The characterisation of molecular boric acid by mass
spectrometry and matrix isolation infrared spectroscopy. J Chem Soc Dalton Trans 1645–
1652
67. Gilson TR (1991) Characterization of ortho- and meta-boric acids in the vapour phase.
J Chem Soc Dalton Trans 2463–2466
68. Andrews L, Burkholder TR (1992) Infrared spectra of molecular B(OH)3 and HOBO in solid
argon. J Chem Phys 97:7203–7210
69. Gupta A, Swanson DK, Tossell JA, Gibbs GV (1981) Calculation of bond distances,
one-electron properties and electron density distributions in first-row tetrahedral hydroxy and
oxyanions. Am Miner 66:601–609
70. Hess AC, McMillan PF, O’Keeffe M (1988) Torsional barriers and force fields in H4TO4
molecules and molecular ions (T = C, B, Al, Si). J Phys Chem 92:1785–1791
71. Nielsen JR, Ward NE (1937) Raman spectrum and structure of the metaborate ion. J Chem
Phys 5:201
72. Edwards JO, Morrison GC, Ross VF, Schultz JW (1955) The structure of the aqueous borate
ion. J Am Chem Soc 77:266–268
73. Oertel RP (1972) Raman study of aqueous monoborate-polyol complexes. Equilibria in the
monoborate-1,2-ethanediol system. Inorg Chem 11:544–549
74. Liu Z, Gao B, Hu M, Li S, Xia S (2003) FT-IR and Raman spectroscopic analysis of hydrated
cesium borates and their saturated aqueous solution. Spectrochim Acta A 59:2741–2745
75. Zhu FY, Fang CH, Fang Y, Zhou YQ, Ge HW, Liu HY (2014) Structure of aqueous
potassium metaborate solution. J Mol Struct 1070:80–85
76. Attina M, Cacace F, Occhiucci G, Ricci A (1992) Gaseous borate and polyborate anions.
Inorg Chem 31:3114–3117
77. Waton G, Mallo P, Candau SJ (1984) Temperature-jump rate study of the chemical relaxation
of aqueous boric acid solutions. J Phys Chem 88:3301–3305
Construction of a Potential Energy Surface
Based on a Diabatic Model for
Proton Transfer in Molecular Pairs
1 Introduction
Rather than examine the adiabatic PES, another approach to chemical reactions is
to analyze the diabatic PES. In contrast to the adiabatic potential, the diabatic
potential presents electronic states that change constantly to confine the eigenstates
of the electronic Hamiltonian. The approaches using diabatic picture have been
utilized various area in chemistry and physics where the coupling between nuclei
and electrons such as vibronic coupling [4–6]. There are some approaches to
describing the diabatic potential [7], constructing using some valence bond
(VB) electronic wave functions [8–11]. Especially, empirical valence bond
(EVB) [12] or multistate empirical valence bond (MS-EVB) [13] approach extended
EVB is used the molecular mechanical functions to construct the PES and applied to
the molecular dynamics (MD) simulations for many proton transfer systems [12, 14–
29]. Furthermore, quantum dynamical approach using molecular mechanical func-
tions (double Morse potential) for proton transfer was also performed [30]. How-
ever, to construct PES using diabatic potentials corresponding to reactant and
product states, which is based on VB picture, is important for understanding
chemical reactions in terms on chemical bond character. Although VB structures are
usually not orthogonal, in this study we consider orthonormal VB structures. To
basic idea, consider a two-state VB electronic wave function as the diabatic basis:
where jϕ1 ⟩ and jϕ2 ⟩ are VB wave functions that describes the electronic structure of
the reactant and product states, respectively. The lowest adiabatic potential energy
Vad is then given by the lower root of the 2 × 2 secular equation; specifically:
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
di
di 2
V di + V22
di
V11 − V22
V ad = 11 − + V12 di 2 , ð2Þ
2 2
where
di
V11 = ⟨ϕ1 jH jϕ1 ⟩, ð3Þ
di
V22 = ⟨ϕ2 jH jϕ2 ⟩, ð4Þ
di
V12 = ⟨ϕ1 jH jϕ2 ⟩. ð5Þ
di di
V11 and V22 are the potential energies for the two VB structures of the reactant
di di di
and product states, respectively. In this approach V11 , V22 and V12 function forms
including parameters can be obtained to fit in experimental or ab initio data. In these
di di
works, V11 and V22 are related to use of molecular mechanics potential functions,
especially, which are taken as the harmonic normal-mode potential or Morse
potential etc. [14, 15, 17, 18, 25, 26]. On the other hand, although the selection of
di di
V12 is less obvious, Gaussian function as V12 proposed by Chang and Millar [14]
Construction of a Potential Energy Surface … 181
has been widely used. However, they has not been confirmed whether the obtained
diabatic potentials produce the reliable adiabatic potentials or not, although these
functions are obtained to fit in some data. Additionally, to obtain these analytical
functions with parameters uniquely or using fewer data is important to construct the
PES for various chemical reactions. Therefore, a proposal of simple method for
di di
light or more uniquely construction of the diabatic potentials (V11 and V22 ) and
di
non-diagonal matrix element (V12 ) using the analytical functions is important to
analyze the chemical reaction and it can be widely applied to describing the large
molecular systems such as proteins.
In this study, we focus on one-dimensional proton transfer models and suggest a
simple construction method of global PES for intermolecular proton transfer by use
di di di
of Morse potential as V11 and V22 and Gaussian function as V12 in the diabatic
potential matrix and confirm the validity to use these potential functions. In addi-
tion, we investigate whether it is possible to apply to the proton transfer for various
intermolecular distance. Here, we focused on the proton-bonded four bimolecular
models for ammonia (AmH+-Am) and imidazole (ImH+-Im) pairs as symmetrical
homo-molecular proton transfer systems and for imidazole-ammonia (ImH+-Am)
and ammonia-water (AmH+-Wat) pairs as asymmetrical hetero-molecular systems,
the four model structures of which were shown in Fig. 1. We investigate the
portability to use the Morse potential and Gaussian function as the diabatic potential
matrix elements by comparison the transformed adiabatic potential from the dia-
batic one with the calculated by DFT calculation in these systems. Finally, we
di di
discuss the proton transfer characters using obtained diabatic potential (V11 and V22 )
di
and non-diagonal elements (V12 ) for homo- and hetero-molecular pairs.
Fig. 1 Four model structures of the proton-bonded a AmH+-Am, b ImH+-Im, c AmH+-Wat, and
d ImH+-Am pairs
182 Y. Hori et al.
In the construction of PES for proton transfer, we focused on the four proton transfer
models: (a) AmH+-Am, (b) ImH+-Im as homo-molecular pairs, and (c) ImH+-Am,
(d) AmH+-Wat as hetero-molecular pairs (Fig. 1). Figure 2 shows also these models
and potential energy coordinates. For homo-molecular pairs, coordinate R denotes
the intermolecular distance and x is the translating proton position that defines as a
displacement from the center of the intermolecular distance. For hetero-molecular
pairs, r is used as the displacement between a proton-bonded atom and the proton.
In this study, we constructed a PES for proton transfer by using diabatic picture.
Diabatic potentials can be constructed using a variety of valence bond (VB) con-
figurations [7–11]. Here, we considered a two-state VB electronic wave function as
the diabatic basis corresponding to the reactant and product states, i.e., Equa-
tion (1). In particular, the EVB approach can be related to use of molecular
mechanical potential functions describing the molecular vibration, which is rational
for understanding chemical reaction. In most cases, the diagonal matrix elements
di di
(V11 and V22 ) are taken as the harmonic normal-mode or Morse potentials [14, 15,
17, 18, 25, 26]. On the other hand, Chang and Millar [14] suggested the use of a
di
generalized Gaussian function as the non-diagonal matrix element (V12 ).
di di
Therefore, we selected the Morse function as the V11 and V22 , and Gaussian
di
function as the V12 to construct PES for proton transfer. For homo-molecular pairs,
di di di
V11 , V22 and V12 are explicitly defined as
+
Fig. 2 Proton transfer model (a) AmH -Am (b) ImH+-Im
and potential energy
coordinates for a AmH+-Am, H x H H x
b ImH+-Im, c AmH+-Wat, H
N H
N H N N H N N
and d ImH+-Am H
H R H R
2
di
V11 ðx; RÞ = D 1 − e − kðx + x0 Þ + c, ð6Þ
2
di
V22 ðx; RÞ = D 1 − ekðx − x0 Þ + c, ð7Þ
di
V12 ðx; RÞ = A exp − bx2 . ð8Þ
di di di
For hetero-molecular pairs, V11 , V22 and V12 are defined as
2
di
V11 ðr; RÞ = D1 1 − e − k1 ðr − r0 Þ + c, ð9Þ
0 2
di
V22 ðr; RÞ = D2 1 − e − k2 ðR − r − r0 Þ + c + D3 , ð10Þ
di
V12 ðr; RÞ = A exp − bðr − rc Þ2 . ð11Þ
di di
The remaining parameters k, k1 , and k2 included in V11 and V22 are considered to
be dependent on the R coordinate. These parameters were estimated by comparison
with the adiabatic potential obtained from DFT calculations according to Eq. (2).
di di
V11 and V22 describes molecular vibration for the reactant and product states,
di di
respectively. Therefore, we assumed that V11 and V22 reproduced the adiabatic
184 Y. Hori et al.
Table 1 Diabatic potential energy parameters for (a) AmH+-Am, (b) ImH+-Im, (c) ImH+-Am,
and (d) AmH+-Wat
(a) AmH+-Am (b) ImH+-Im
−1
D(kJ mol ) r0 (Å) D(kJ mol−1) r0 (Å)
664.758 1.027 668.81 1.0148
(c) ImH+-Am
D1 (kJ mol−1) D2 (kJ mol−1) r0 (Å) 0
r0 (Å)
668.81 664.758 1.015 1.027
(d) AmH+-Wat
D1 (kJ mol−1) D2 (kJ mol−1) r0 (Å) 0
r0 (Å)
725.69 593.731 1.027 0.977
potential, when the proton position was closer to a binding atom than a stable point
di
(r < r0 ). Here, we estimated the parameter k by comparing them with the V22 and
adiabatic potential at x = 2x0 for homo-molecular pairs for fixed R. In the same way,
for hetero-molecular pairs, k1 and k2 were determined by comparing with the
0
adiabatic potential at r = 2r0 − R ̸2 ≡ rk1 and r = 3R ̸2 − 2r0 ≡ rk2 , respectively, for
fixed R. These relations are explicitly defined as
di
V22 ð2x0 Þ = V ad ð2x0 Þ, ð13Þ
di
V11 ðrk1 Þ = V ad ðrk1 Þ, ð14Þ
di
V22 ðrk2 Þ = V ad ðrk2 Þ. ð15Þ
0sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi1
1 V ad ðrk2 Þ − c − D3 A
k2 = 0 ln@1 + . ð18Þ
r k2 + r 0 − R D2
di
Subsequently, parameters A and b of V12 depending on the coordinate R were
estimated. These parameters were also estimated by comparison with the adiabatic
potential obtained from DFT calculations and using obtained diabatic potentials
di di
(V11 and V22 ). Determination of these parameters was conducted according to the
following steps for each pairs.
Construction of a Potential Energy Surface … 185
First, parameter A was estimated. Because this parameter represents the ampli-
tude of the Gaussian type function, parameter b vanishes at x = 0 for
homo-molecular pairs. Thus, the relationship between the adiabatic and diabatic
potentials at x = 0 and fixed R is given by
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
di di ð0Þ 2
di
V11 ð0Þ + V22
di
ð 0Þ V11 ð0Þ − V22
− + V12 di 2 ð0Þ = V ad ð0Þ. ð19Þ
2 2
di di
Parameter A was then estimated by using obtained V11 and V22 , which was
expressed as
di
A = V11 ð0Þ − V ad ð0Þ, ð20Þ
x0 x0 sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ffi
di x0
di x0 2 x x
di
V11 di
+ V22 V11 2 − V22 0 0
2 2
− 2
+ V12 di 2 = V ad . ð22Þ
2 2 2 2
To obtain the adiabatic PES V ad ðx, RÞ or V ad ðr, RÞ for the proton transfer reaction,
quantum chemical calculations were performed with stepwise movement of the
proton and intermolecular distances by 0.02 and 0.05 Å, respectively. The
geometries of the molecules were then kept in the minimum energy structures,
except for the proton position (x or r) and the intermolecular distance (R). Minimum
energy structures, as shown in Fig. 1, were determined by geometrical optimiza-
tion. Finally, the adiabatic potential energies were fitted using a polynomial series
function of the fourth order with respect to x or r and of the third order with respect
to R [31, 32]. In previous work [33, 34], the proton transfer of ImH+-Im systems
was discussed with the B3LYP approach. Therefore, all DFT calculations were
performed at the B3LYP/aug-cc-pVDZ level using the Gaussian-09 package [35].
(a) AmH+-Am
R=2.70 R=3.05
300 300
Energy (kJ/mol)
Energy (kJ/mol)
200 200
100 100
0 0
-0.6 -0.4 -0.2 0 0.2 0.4 0.6 -0.7 -0.5 -0.3 -0.1 0.1 0.3 0.5 0.7
x( ) x( )
(b) ImH+-Im
R=2.66 R=3.01
300 300
Energy (kJ/mol)
Energy (kJ/mol)
200 200
100 100
0 0
-0.6 -0.4 -0.2 0 0.2 0.4 0.6 -0.8 -0.4 0 0.4 0.8
x( ) x( )
(c) ImH+-Am
R=2.77 R=3.12
200 300
Energy (kJ/mol)
Energy (kJ/mol)
250
150
200
100 150
100
50
50
0 0
0.8 1 1.2 1.4 1.6 1.8 2 0.8 1.2 1.6 2 2.4
r( ) r( )
(d) AmH+-Wat
R=2.69 R=3.04
400 400
Energy (kJ/mol)
Energy (kJ/mol)
300 300
200 200
100 100
0
0.8 1 1.2 1.4 1.6 1.8 2 0
0.8 1.2 1.6 2 2.4
r( ) r( )
di di
Fig. 3 Diagonal matrix elements, V11 (blue line) and V22 (red line), and non-diagonal matrix
di
element V12 (yellow line) in diabatic potential matrix computed using obtained potential
parameters at some intermolecular distance R for a AmH+-Am, b ImH+-Im, c ImH+-Am, and
d AmH+-Wat
188 Y. Hori et al.
Energy (kJ/mol)
Energy (kJ/mol)
140 140
100 100
60 60
20 20
-20 -20
-0.6 -0.4 -0.2 0 0.2 0.4 0.6 -0.6 -0.4 -0.2 0 0.2 0.4 0.6
x( ) x( )
R=2.85 R=3.05
180 180
Energy (kJ/mol)
Energy (kJ/mol)
140 140
100 100
60 60
20 20
-20 -20
-0.6 -0.4 -0.2 0 0.2 0.4 0.6 -0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8
x( ) x( )
180
Energy (kJ/mol)
180
130
140
80 100
60
30
20
-20 -20
-0.5 -0.3 -0.1 0.1 0.3 0.5 -0.6 -0.4 -0.2 0 0.2 0.4 0.6
x( ) x( )
R=2.81 R=3.01
180 220
Energy (kJ/mol)
Energy (kJ/mol)
140 180
100 140
100
60
60
20
20
-20 -20
-0.6 -0.4 -0.2 0 0.2 0.4 0.6 -0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8
x( ) x( )
Fig. 4 Transformed adiabatic potential derived from the diabatic potentials using Eq. (2) (red
line) and from DFT (blue dots) at some intermolecular distance R for a AmH+-Am and b ImH+-Im
diabatic potential employed four reference points (x = x0 , 2x0 , 0 and x0 ̸2) at fixed
R for homo-molecular pairs, while five reference points (r = r0 , rk1 , rk2 , rc and rb )
for hetero-molecular pairs. Thus, the whole potentials of proton transfer are
described by using energies of 40 or 50 reference points, while the number of DFT
data points to describe the whole adiabatic potential required approximately 500
points. Therefore, the PES describing the entire proton transfer system for diabatic
picture can be obtained using less than one-tenth of the reference points required for
Construction of a Potential Energy Surface … 189
Energy (kJ/mol)
160
Energy (kJ/mol)
160
120 120
80 80
40 40
0 0
-40 -40
0.8 1 1.2 1.4 1.6 1.8 2 0.8 1 1.2 1.4 1.6 1.8 2 2.2
r( ) r( )
R=2.92 R=3.12
200 200
Energy (kJ/mol)
160
Energy (kJ/mol)
160
120 120
80 80
40 40
0 0
-40 -40
0.8 1 1.2 1.4 1.6 1.8 2 2.2 0.8 1 1.2 1.4 1.6 1.8 2 2.2 2.4
r( ) r( )
450
350
350
250
250
150 150
50 50
-50 -50
0.8 1 1.2 1.4 1.6 1.8 2 0.8 1 1.2 1.4 1.6 1.8 2 2.2
r( ) r( )
R=2.84 R=3.04
450 450
Energy (kJ/mol)
Energy (kJ/mol)
350 350
250 250
150 150
50 50
-50 -50
0.8 1 1.2 1.4 1.6 1.8 2 2.2 0.8 1.2 1.6 2 2.4
r( ) r( )
Fig. 5 Transformed adiabatic potential derived from the diabatic potentials using Eq. (2) (red
line) and from DFT (blue dots) at some intermolecular distance R for c ImH+-Am and d AmH+-
Wat
Table 2 Coefficient of determination (R-squared) between adiabatic potential derived from the
diabatic potentials and from DFT
(a) AmH+-Am (b) ImH+-Im (c) ImH+-Am (d) AmH+-Wat
R-squared 0.966 0.974 0.932 0.965
190 Y. Hori et al.
-2)
Wat 7.2 (d) AmH-Wat
b(
5.2
3.2
1.2
2.5 2.6 2.7 2.8 2.9 3 3.1 3.2
R( )
the adiabatic picture. Furthermore, our method can be applied to proton transfer
system even when the transition state (TS) cannot be calculated, although the
information of TS requires the construction of potential in the previous work.
Therefore, it is concluded that our procedures are useful for PES construction by
using the diabatic picture for proton transfer systems and can be applied to large
molecular systems such as proteins.
Finally, we discuss the obtained potential parameters. Figure 6 shows that
intermolecular dependence of potential parameter b of V12 di
for (a) AmH+-Am,
+ + +
(b) ImH -Im, (c) ImH -Am, and (d) AmH -Wat. The values of parameter b, which
describe the spread of the Gaussian function, decreased as R increased. This result
di
indicates that the non-diagonal matrix element (V12 ) is broadly distributed along the
proton transfer coordinate and the bond mixture between the reactant and the
di
product states occurs over a wide range, not only at the TS. In addition, because V12
is broadly distributed along the intermolecular distance, the proton can be formed
mixture between reactant and product states and transferred at the location formed
di
hydrogen bond. To clarify the effect of non-diagonal matrix element V12 , the ratio
di
of amplitude for the V12 (parameter A) divided by the crossing point energy of the
di di
V11 and V22 was estimated, the results of which were shown in Fig. 7. According to
0
2.5 2.6 2.7 2.8 2.9 3 3.1 3.2
R( )
Construction of a Potential Energy Surface … 191
di
Eq. (2), V12 contributes to the stability of the adiabatic potential V ad and the
A ̸V11
di
ðx = 0 or r = rc Þ is assumed to be determined the rate of contribution to the
possibility of proton transfer. For both homo-molecular pairs, the value of A ̸ V11 di
4 Conclusion
References
Abstract The possible stable geometries of the atmospheric negative core ion
NO3− ðHNO3 Þ2 and its monohydrate were theoretically investigated with the second
order Møller-Plesset perturbation theory (MP2) in consideration of the effect of
electron correlation. For both ionic clusters, we obtained the different stable
geometries from the previous study by Drenck and coworkers (Int J Mass Spectrom
273:126–131, 2008) [1] with the density functional theory of Becke 3-parameters
hybrid functional (B3LYP). The non-planar geometry with two hydrogen-bondings
between one oxygen atom on NO3− and each hydrogen atom of two HNO3 frag-
ments is found as the most stable structure of the core ion at 0 K. For the mono-
hydrate, the most stable geometry at 0 K is found as the H2O-embedded form in
which one water molecule is located at the center of the cluster with
hydrogen-bondings to NO3− and HNO3 fragments. Our results show that the
hydrogen bond network of the core ion can be strongly perturbed by a single water
molecule. We also discussed the relative abundance of conformers of these ionic
clusters under a finite temperature.
1 Introduction
[6]. Experimental results observed in situ mass spectrometry show the existence of
atmospheric negative ion NO3− ðHNO3 Þ2 in the upper troposphere (altitude between
9 and 12 km) [7, 8]. This negative ion is also known as a dominant negative ion at
the upper troposphere region [7].
Recently, Sekimoto and Takayama established the atmospheric pressure corona
discharge ionization (APCDI) technique, which enables us to reproducibly generate
negative ions [9]. They reported the existence of stable negative ion water clusters,
NO3− ðHNO3 Þ2 ðH2 OÞn , and the specific stability referred to as a magic number for
n = 8 [10]. Geometric structures of these water clusters are, however, still unclear
even for the core ion and its monohydrate (n = 1).
From theoretical points of view, Drenck and coworkers reported stable structures
of NO3− ðHNO3 Þ2 ðH2 OÞn (n = 0–4) obtained at B3LYP/6-31++G** level of
density functional theory (DFT) calculations [1]. They also reported the stable
structures of NO3− ðHNO3 Þm ðH2 OÞn up to n + m = 6, and assessed the validity of
theoretical calculations by comparing theoretical dissociation energies of the
NO3− ðHNO3 Þm ðH2 OÞn cluster into NO3− , mHNO3, and nH2O fragments to the
experimental results with mass-analyzed ion kinetic energy (MIKE) spectra mea-
surement [11]. Their theoretical results agree with the experiments reasonably, but
assumed only one kind of geometry for each NO3− ðHNO3 Þm ðH2 OÞn clusters despite
that a water cluster should generally have various kinds of conformers. In order to
elucidate the stable geometries of NO3− ðHNO3 Þ2 and its hydrates, a more com-
prehensive geometry searching with first-principles calculations must be
indispensable.
In this study, to elucidate stable geometries of these ionic clusters in details, we
theoretically analyzed stable geometries of the negative core ion NO3− ðHNO3 Þ2 and
its monohydrate in consideration of a lot of possible conformers with the post
Hartree-Fock ab initio method. We also discussed the relative abundance of con-
formers of these ionic clusters under a finite temperature.
2 Computational Details
We employed the second order Møller-Plesset perturbation theory (MP2) with 6-31
++G** Gaussian type basis sets in ab initio calculations of the negative core ion,
and its monohydrate. The basis set superposition error (BSSE) is not corrected
because of the less convergence in BSSE corrected geometrical optimization pro-
cedure. The harmonic approximation was used to evaluate the zero-point vibration
energy (ZPE) and Gibbs free energy. Natural Population Analysis (NPA) [12] was
used to analyze electronic populations on each atom. All calculations were per-
formed with GAUSSIAN 09 program package [13].
In the comformational searching, we picked up the initial geometries to be a
molecular cluster consisting of one NO3− , two HNO3’s, and one H2O fragments,
and optimized all the geometric degrees of freedom of the cluster simultaneously.
Ab Initio Investigations of Stable Geometries … 195
δ H = +0.49
δ N = +0.66 δ N = +0.68
δ H = +0.51
δ O = -0.34 δ O = -0.98
Fig. 1 Geometries of molecular fragments, NO3− , HNO3 , and H2O obtained with MP2/6-31+
+G** calculations. The δX means the NPA charge on the element X
Table 1 Intermolecular interaction energies with zero-point vibration correction (EInt, unit in
kcal/mol) between two fragments obtained with MP2/6-31++G** calculation. The EInt value for
the complex X⋯Y is calculated as EInt = EZPE(X) + EZPE(Y) − EZPE(X⋯Y), where EZPE(X) is
the sum of electronic total energy and the zero-point energy (ZPE) of the system X
Complex EInt Exptl. [11]
NO3− ⋯H2 O 14.9 14.6 ± 0.2
NO3− ⋯HNO3 30.6
HNO3 ⋯HNO3 7.9
HNO3 ⋯H2 O 9.2
The stable structures of these molecular fragments are shown in Fig. 1. The ener-
getic stabilities of NO3− ðHNO3 Þ2 and its monohydrate can be mainly determined by
the O⋯H − O type intermolecular hydrogen-bonding (HB) between the fragments,
where the ionic HB between NO3− and HNO3 fragments is the most strong inter-
action among possible two fragments as shown in Table 1. We note here that MP2/
6-31 ++G** level of ab initio calculations reasonably reproduce the corresponding
experimental intermolecular interaction energy between NO3− and H2O with MIKE
spectra [11]. In the geometry optimizations of NO3− ðHNO3 Þ2 , initial geometries
were chosen to have HBs between the fragments as much as possible, where the
total number of initial geometries is 45 including conformers having a different
angle between molecular planes of NO3− and HNO3. In the case of the monohy-
drate, all the possible 218 hydrogen-bonded structures between NO3− , HNO3, and
H2O fragments were chosen as initial geometries.
196 A. Ueda et al.
all
pi ðTÞ = expf − ΔGi ̸kB T g ̸ Z , Z = ∑ exp − ΔGj ̸kB T , ð1Þ
j
where ΔGi is the relative Gibbs energy including the enthalpic (H: the sum of the
electronic energy, ZPE, thermal corrections, etc.) and the entropic terms (−TS). In
Fig. 3, the temperature dependences of pi values for all conformers are shown in the
upper figure (a), and those of the sum of pi values over conformers in each group
are given in the lower figure (b). Figure 3 clearly indicates that the relative abun-
dance of NO3− ðHNO3 Þ2 conformers depends on the temperature, and the inversion
of the total abundance of the groups α and β is found at around 250 K. We note
here that the most stable structure α1 at 0 K is a dominant conformer in the low
temperature region below 90 K, while the structure of β9 has the largest relative
abundance in high temperature region above 350 K due to the large entropic
contribution from the lowest twisting mode of two HNO3 fragments (the harmonic
vibrational frequency ωe = 10 cm−1). Such temperature dependence of the relative
abundances as well as the inversion of the abundance is arising from the entropic
Ab Initio Investigations of Stable Geometries … 197
Fig. 2 Stable geometries of NO3− ðHNO3 Þ2 obtained with MP2/6-31++G** calculations. The
definitions of each group are given in the text. The ΔEZPE (unit in kcal/mol) means the relative
energy with zero-point vibration correction from the most stable structure α1
contributions and the difference in the number of conformers in each group (nu-
merical contribution). Since the conformer α2 has the largest pi value in the tem-
perature range from 90 to 350 K as shown in Fig. 3a, at around the inversion
temperature, the numerical advantage of the group β conformers plays a dominant
role in the inversion of the relative abundance.
198 A. Ueda et al.
(a)
(b)
(a) (b)
(c) (d)
Fig. 4 The most stable structures of NO3− ðHNO3 Þ2 H2 O in the group A–D obtained with MP2/
6-31 ++G** calculations. The definitions of each group are given in the text. The ΔEZPE (unit in
kcal/mol) means the relative energy with zero-point vibration correction from the most stable
structure. The value in the parenthesis is the maximum ΔEZPE value in each group
Since the relative energy of most conformers are within a few kcal/mol (ΔEZPE
values of 76 conformers (84%) are within 3 kcal/mol), the relative abundances of
conformers should depend on a thermal condition as in the case of the core ion.
Figure 5 shows the theoretical temperature dependence of the relative abundances
of conformers: Fig. 5a the temperature dependences of pi values for all conformers,
and Fig. 5b those of the sum of pi values over conformers in each group. As shown
in Fig. 5b, the H2O-embedded conformers of the group C (red solid line) have the
largest relative abundance among all groups in the low-temperature region below
330 K. For the H2O-attached conformers, the conformers of the group A (blue
dashed lines) have a small relative abundance over whole temperature range, but the
group B conformer (green dotted line) has the largest relative abundance above
330 K. The relative abundance of OH − ðHNO3 Þ3 conformer is less than 1% in this
(a)
(b)
temperature range. As shown in the inset figure in Fig. 5b, the inversion of the
relative abundance of the H2O-embedded and attached conformers occurs at around
270 K. At around this inversion temperature, the numerical advantage of H2O-at-
tached conformers contributes to the inversion dominantly, because the relative
abundances of H2O-embedded conformers are larger than that of H2O-attached
conformers. The large relative abundances of H2O-attached conformers in the high
temperature region are due to the entropic contributions from slow molecular
vibrations relevant to the attached H2O molecule ranging from 20 to 30 cm−1, e.g.
H2O rocking mode.
4 Conclusion
Acknowledgements The present study was supported by Grant-in-Aid for Scientific Research
and for Priority Areas by Ministry of Education, Culture, Sports, Science and Technology, Japan
(KAKENHI). A part of the present computations were performed using Research Center for
Computational Science, Okazaki, Japan.
References
1. Drenck K, Hvelplund P, Nielsen SB, Panja S, Støchkel K (2008) Int J Mass Spectrom
273:126–131
2. Yu F, Turco RP (2000) Geophys Res Lett 27:883–886
3. Fend J, Möller D (2004) J Atmos Chem 48:217–233
4. Harrison RG, Carslaw KS (2003) Rev Geophys 41:1012–1027
5. Singh A, Agrawal M (2008) J Environ Biol 29:15–24
202 A. Ueda et al.
Taku Onishi
Keywords Helium ⋅
Hydrogen ⋅ Molecular orbital ⋅ Covalent bonding
Chemical bonding rule
1 Introduction
It has been recognized that helium cannot form covalent bonding with other atoms
and molecules. It is because helium has the stable closed shell configuration.
Recently, Helgaker et al. demonstrated that helium clusters are not dispersed under
the strong magnetic field [1]. Several types of helium clusters such as He3, He4 and
He6 were optimized under the circumstance [2]. On the other hand, without mag-
netic field, many quantum chemical calculations were performed, from the interest
of van der Waals interaction [3–5] and potential energy [6–8]. However, molecular
orbital analysis was not performed for helium clusters and helium-including
clusters. In our previous study, the chemical bonding character of helium dimer
T. Onishi (✉)
Center of Ultimate Technology on Nano-electronics, Mie University, Mie, Japan
e-mail: [email protected]; [email protected]
T. Onishi
Centre for Theoretical and Computational Chemistry (CTCC), Department of Chemistry,
University of Oslo, Oslo, Norway
T. Onishi
Department of Applied Physics, Osaka University, Osaka, Japan
(He–He) [9, 10] was investigated by coupled cluster calculations. At a local min-
imum, bonding and anti-bonding molecular orbitals (MOs) were obtained. By the
use of our chemical bonding rule [11, 12], it was concluded that covalent bonding is
formed at optimized structure, due to orbital overlap between helium 1s orbitals.
Bond order is approximate equation to investigate a stability of chemical
bonding in two-atom system:
Na − Nb
N=
2
where Na and Nb denote the number of electrons in bonding and its anti-bonding
MOs. In the case of helium dimer, bonding and anti-bonding MOs are formed
between helium 1s orbitals. Since two electrons occupy both MOs (Na = Nb = 2),
N becomes zero. It was understood that covalent bonding is formed, even if bond
order is zero [10].
In this study, we have investigated chemical bonding character between helium
and hydrogen. The coupled cluster calculations have been performed for He–H
model. Here, three different hydrogen formal charges have been assumed: (1) pos-
itive (H+); (2) neutral (H); (3) negative (H−).
2 Computations
The calculations presented here were performed using the coupled cluster singles
and doubles (CCSD) method. It is because CCSD method accurately reproduces
accurate interatomic distance, electronic state and molecular spectra for simple
molecule [13, 14]. We used aug-cc-pVTZ basis set for helium and hydrogen [15].
All calculations were performed with the Gaussian program [16]. To investigate
chemical bonding character between helium and hydrogen, the simple He–H model
was constructed. Three types of He–H+, He–H and He–H− models were constructed.
Potential energy curves were obtained, changing the interatomic distance. In addi-
tion, zero point vibrational energy has been also obtained at the optimized structure.
In order to judge chemical bonding character such as covalency and ionicity in the
calculated MOs, chemical bonding rule is very useful and applicable (see Fig. 1).
How to utilize chemical bonding rule
1. In MOs including outer shell electrons, check whether the orbital overlap
between helium and hydrogen exists or not.
A Theoretical Study of Covalent Bonding Formation … 205
Figure 2 shows potential energy curve of He–H+ model, changing the interatomic
He–H distance. A local minimum is given at 0.776 Å. Two electrons occupy MO1,
which is only one occupied MO. The wave-function of MO1 at a local minimum is
-1790
-1800
-1810
Total Energy [kcal/mol]
-1820
-1830
-1840
-1850
-1860
-1870
-1880
0.5 1.0 1.5 2.0 2.5 3.0
He-H distance [Å]
Fig. 2 Potential energy curve of He–H+ model, changing the interatomic He–H distance
206 T. Onishi
e
1s orbital 1s orbital
MO1
ψ MO1 ðHe − H + Þ = 0.13 ϕHð1s′ Þ + 0.12 ϕHð1s′′ Þ + 0.35 ϕHeð1s′ Þ + 0.45 ϕHeð1s′′ Þ + 0.16 ϕHeð1s′′′ Þ
There is orbital overlap between helium 1s orbital and hydrogen 1s orbital. From
chemical bonding rule, it is concluded that helium forms covalent bonding with
hydrogen. Mulliken charge densities of helium and hydrogen are 0.315 and 0.685,
respectively.
Figure 3 depicts the schematic drawing of electrons and orbitals in He–H+
model. Helium 1s electrons are shared by both helium and hydrogen, through
covalent bonding formation between helium and hydrogen. Hence, the interatomic
He–H+ distance becomes smaller.
Figure 4 shows potential energy curve of He–H model, changing the interatomic
He–H distance. A local minimum is given at 3.577 Å. It is much larger than the
interatomic He–H+ distance (0.776 Å). Alpha and beta electrons are occupied in
different MO1α and MO1β, respectively. The wave-function of MO1α is
It is found that MO1α and MO1β are paired. Orbital energies of MO1α and
MO1β are slightly different: −0.91792 au for MO1α; −0.91787 au for MO1β. It is
due to broken spin symmetry. MO2α has no paired beta MO. The wave-function of
MO2α is
A Theoretical Study of Covalent Bonding Formation … 207
-2133.770
-2133.775
-2133.780
Total Energy [kcal/mol]
-2133.785
-2133.790
-2133.795
-2133.800
-2133.805
-2133.810
-2133.815
3.0 3.2 3.4 3.6 3.8 4.0 4.2 4.4 4.6 4.8 5.0
He-H distance [Å]
Fig. 4 Potential energy curve of He–H model, changing the interatomic He–H distance
There is no orbital overlap between helium and neutral hydrogen. From chemical
bonding rule, no covalent bonding is formed in He–H model. Mulliken spin den-
sities of helium and hydrogen are 0.00 and 1.00, respectively. MO2α is responsible
for the spin density.
Figure 5 depicts the schematic drawing of electrons and orbitals in He–H model.
Electron repulsion between two helium 1s electrons and hydrogen 1s electron is
dominative, though Coulomb interaction exists between helium 1s electrons and
positive hydrogen atomic nucleus. It is considered that the repulsion lets the
interatomic He–H distance elongated.
Figure 6 shows potential energy curve of He–H− model, changing the interatomic
He–H distance. A local minimum is given at 6.452 Å. It is much larger than He–H+
and He–H models. It is found that He–H− is weakly bounded. Four electrons
occupy MO1 and MO2. The wave-function of MO1 is
He e H
e
1s orbital 1s orbital
MO2(α)
MO1(β)
MO1(α)
-2150.555
-2150.560
-2150.565
Total Energy [kcal/mol]
-2150.570
-2150.575
-2150.580
-2150.585
-2150.590
-2150.595
-2150.600
5.0 5.5 6.0 6.5 7.0 7.5 8.0
He-H distance [Å]
Fig. 6 Potential energy curve of He–H− model, changing the interatomic He–H distance
ψ MO2 ðHe − H − Þ = 0.16 ϕHð1s′ Þ + 0.27 ϕHð1s′′ Þ + 0.41 ϕHð1s′′′ Þ + 0.37 ϕHð2s′ Þ
e e
He H-
e e
1s orbital 1s orbital
MO2
MO1
1s orbital but also hydrogen 2s orbital. Mulliken charge densities of helium and
hydrogen are 0.00 and –1.00, respectively.
Figure 7 depicts the schematic drawing of electrons and orbitals in He–H−
model. There is electron repulsion between two helium 1s electrons and two
hydrogen 1s electrons, though Coulomb interaction exists between helium 1s
electrons and positive hydrogen atomic nucleus. It is considered that this effect is
stronger than He–H− model, due to existence of two hydrogen 1s electrons.
The dissociation energies for He–H+, He–H and He–H− models can be estimated
from the total energy difference between local minimum and completely dissociated
point. They were 46.9 kcal/mol, 0.0117 kcal/mol and 0.0191 kcal/mol for He–H+,
He–H and He–H− models, respectively. Except He–H+ model, the values are very
small. It is because the stable covalent bonding between helium and hydrogen is
formed only in He–H+ model. In order to investigate the effect of quantum
vibration on the dissociation, we obtain zero point vibration energy. When zero
point vibration energy is smaller than dissociation energy, hydrogen is kept fixed at
optimized position. On the other hand, when it is larger, the dissociation may be
caused by the external factor.
210 T. Onishi
Zero point vibration energies for He–H+, He–H and He–H− moleds, were 4.58,
0.0467 and 0.0208 kcal/mol, respectively. In He–H+ model, it is much smaller than
dissociation energy. It shows that optimized structure is much stabilized, because
hydrogen is kept fixed at optimized structure. It is concluded that helium can be
strongly bounded with positive hydrogen. On the other hand, zero point vibration
energies are larger than dissociation energies in He–H and He–H− models. In
addition, the dissociation energies are very small and no covalent bonding is
formed.
4 Conclusions
From chemical bonding rule, it was found that helium 1s orbital and hydrogen 1s
orbital forms the stable covalent bonding in He–H+ model. Zero point vibration
energy was estimated to be 4.58 kcal/mol. It is much smaller than dissociation
energy (46.9 kcal/mol). On the other hand, no covalent bonding is formed, and the
small dissociation energy is given in He–H and He–H− models. Helium is weakly
bound with hydrogen.
References
1. Lange KK, Tellgren EI, Hoffmann MR, Helgaker T (2012) Science 337:327
2. Tellgren EI, Reine SS, Helgaker T (2012) Phys Chem Chem Phys 14:9492
3. Tao J, Perdew JP (2005) J Chem Phys 122:114102
4. Zhao Y, Truhlar DG (2006) J Phys Chem A 110:5121
5. Kamiya M, Tsuneda T, Hirao K (2002) J Chem Phys 117:6010
6. Lotrich VF, Bartrett RJ, Grabowski I (2005) Chem Phys Lett 405:43
7. Snook I, Per MC, Russo SP (2008) J Chem Phys 129:164109
8. Allen MJ, Tozer DJ (2002) J Chem Phys 117:11113
9. Onishi T (2016) J Chin Chem Soc 63:83
10. Onishi T (2016) AIP Conf Proc 1790:02002
11. Onishi T (2012) Adv Quant Chem 64:31
12. Onishi T (2015) Adv Quant Chem 70:31
13. Helgaker T, Jorgensen P, Olsen J (2000) Molecular electronic-structure theory. Wiley, p 648
14. Bartlett R, Musial M (2007) Rev Mod Phys 79:291
15. Woon DE, Dunning TH Jr (1994) J Chem Phys 100:2975
16. Gaussian 09, Frisch MJ, Trucks GW, Schlegel HB, Scuseria GE, Robb MA, Cheeseman JR,
Scalmani G, Barone V, Mennucci B, Petersson GA, Nakatsuji H, Caricato M, Li X,
Hratchian HP, Izmaylov AF, Bloino J, Zheng G, Sonnenberg JL, Hada M, Ehara M,
Toyota K, Fukuda R, Hasegawa J, Ishida M, Nakajima T, Honda Y, Kitao O, Nakai H,
Vreven T, Montgomery JA Jr, Peralta JE, Ogliaro F, Bearpark M, Heyd JJ, Brothers E,
A Theoretical Study of Covalent Bonding Formation … 211
Kudin KN, Staroverov VN, Kobayashi R, Normand J, Raghavachari K, Rendell A, Burant JC,
Iyengar SS, Tomasi J, Cossi M, Rega N, Millam JM, Klene M, Knox JE, Cross JB, Bakken V,
Adamo C, Jaramillo J, Gomperts R, Stratmann RE, Yazyev O, Austin AJ, Cammi R,
Pomelli C, Ochterski JW, Martin RL, Morokuma K, Zakrzewski VG, Voth GA, Salvador P,
Dannenberg JJ, Dapprich S, Daniels AD, Farkas Ö, Foresman JB, Ortiz JV, Cioslowski J,
Fox DJ (2009) Gaussian, Inc., Wallingford CT
Small Rhodium Clusters:
A HF and DFT Study–III
Abstract Small neutral and ionic Rhodium clusters Rhn (n = 6, 8, 13) are investi-
gated by ab initio molecular orbital calculations with full optimization at the
Restricted Open Shell Hartree-Fock (ROHF) level with a LANL2DZ basis set, and
with the methods based on Density Functional Theory, B3LYP/MWB, B3LYP/PBE.
The clusters are found favor close-packed icosahedron structures in contrast to pre-
vious theoretical predictions that rhodium clusters should favor cubic motifs. A range
of spin multiplicities are investigated for each cluster and we present the minimum
energy conformation along with the vertical and adiabatic ionization potentials.
1 Introduction
M. A. Mora (✉)
Depto. de Química, Universidad Autónoma Metropolitana, campus Iztapalapa,
Av. Sn. Rafael Atlixco 186, 09340 Mexico, D. F., Mexico
e-mail: [email protected]
M. A. Mora-Ramírez
Depto. Fisicomatemáticas, Facultad de C. Químicas, Benemérita Universidad
Autónoma de Puebla, Sn. Claudio y Sur 22 Col. Sn. Manuel, 72570 Puebla, Mexico
e-mail: [email protected]
several chemical processes such as manufacturing of certain silicon rubbers [7] and
the reduction of benzene to cyclohexane [8]. Rhodium also finds use in the jewelry
industry and as an agent for hardening and improving corrosion resistance [9].
The main problem to obtain a deep atomic-level understanding of the physical
and chemical properties of clusters relies on an accurate determination of their
equilibrium atomic structure, which is not as simple as it might appear. A direct
identification of the equilibrium atomic structure by experimental techniques is very
difficult and only indirect measurements can provide few clues about the atomic
structure. Thus, the combinations of experimental techniques with first-principles
calculations have been used. For example, vibrational spectroscopies combined
with theoretical calculations have lead to important insights into the atomic struc-
ture of small Rh clusters [10, 11]. Isolated metal clusters have also been investi-
gated by Stern-Gerlach molecular-beam deflection experiments [2, 12–17].
However, there are difficulties in the direct identification of the atomic structure
of clusters by experimental techniques. Thus, most of the structural studies have
been based on theoretical calculations, which can directly determine the atomic
structure of clusters using several well-defined algorithms. Several calculations
based on density functional theory (DFT) have focused on these clusters, for
example, on metal particles containing 13 atoms, Rh13 [18–29]. Furthermore, it is
important to mention that few studies have focused on the search for the
lowest-energy structures with most studies assuming predefined structures.
Sophisticated algorithms have been employed in the search for the lowest-energy
structure, namely, generic algorithms (GA) [30], basin-hopping Monte Carlo
(BHMC) [31–34], Monte Carlo (MC) [35], conformational space annealing [29],
taboo search in descriptor space (TSDE) [23, 36], high-temperature molecular
Dynamics (high-T-MD) [22]. Almost all the studies with these algorithms have
been used in combination with empirical pair potentials. These potentials have
difficulties in providing a correct description of the atomic structure [37–39], and
hence, the ground state structures might not be correct.
In this paper, we present HF and DFT calculations on clusters with 6 and 8
rhodium atoms for comparison with theoretical and experimental results. We also
present results for the 13-atom cluster since it is one of the most studied clusters and
to the best of our knowledge there are no experimental results.
It is well known that the method of calculation and the chosen basis set are the two
most important factors in determining the accuracy of results. Ab initio methods
must represent all the electrons in some manner. However, for heavy atoms it is
desirable to reduce the amount of computation burden. One way to do this is by
replacing the core electrons and their basis functions in the wave function by a
potential term in the Hamiltonian. These are called core potentials, effective core
potentials (ECP) or relativistic effective core potentials (RECP). In this work, we
Small Rhodium Clusters: A HF and DFT Study–III 215
use the ECP LANL2DZ [40–42] potential, which is one of the most widely used for
heavy elements.
Another factor to consider in selecting the method of calculation is the spin
contamination. A high spin contamination can affect the geometry and population
analysis and significantly affect the spin density. The error introduced by spin
contamination is unacceptable when systems with transition metals are investigated.
The restricted open-shell Hartree-Fock (ROHF) ab initio method is one of the best
ways to avoid spin contamination and obtain a reliable wave function.
We performed ab initio molecular orbital calculations with full optimization at
the ROHF level with a LANL2DZ basis set. Also in this research we employed the
B3LYP hybrid functional with the small-core quasirelativistic approach of Wood
and Boring [43] MWB ECP, and, the General Gradient Approximation
(GGA) formulated by Perdew, Burcke, and Hernzerhof (PBE) [44, 45] since it has
been shown that these hybrid functional can yield reliable energetics and structural
results for other metal compounds. The geometries were adjusted until a stationary
point on the potential energy surface was found, using the Berny algorithm [46, 47]
for the minimization.
Figure 1 shows the Rh6 optimized geometry calculated with the B3LYP/MWB
chemical model obtained from different initial geometries; Fig. 1a shows the final
geometry obtained from an initial octahedron, Fig. 1b is a final triangular prism
conformer, Fig. 1c is a deformed pentagonal pyramid which converges to an dis-
torted trigonal prism or capped tetrahedron as in Fig. 1d.
Table 1 shows the relative energy (eV) of the Rh6 cluster obtained with the
B3LYP/MWB chemical model; we searched the spin multiplicity up to 23. The
lowest energy isomer is an octahedron with a multiplicity equal to 9, followed by
the distorted pentagonal pyramid with a multiplicity of 11. Finally the trigonal
prism isomer has M = 9. The minimum energy isomer is in complete agreement
with the experimental results [10, 48], and with previous theoretical results [21, 24,
25, 48–56]. The relative energy of ROHF-isomers with different geometry is 1.039,
3.973 eV for pentagonal pyramid and octahedron respectively. For the
B3LYP-isomers with different energy the relative energy is 0.218 for pentagonal
pyramid, and 0.490 eV for trigonal prism. With the ROHF/LANL2DZ method, the
triangular prism has a minimum energy conformation with multiplicity equal to 15.
The more stable pentagonal pyramid is when M = 17, and the square bi-pyramid
with M = 11. A geometry formed by two perpendicular squares (an incomplete
cube) with M = 15 is the fundamental state for this particular geometry. Among
these four geometries, the minimum energy is the triangular prism. All calculations
in this series were performed with the ROHF/LANL2DZ methods and with mul-
tiplicities from 1 to 23 and full optimization.
216 M. A. Mora and M. A. Mora-Ramírez
Fig. 1 Isomers of Rh6 geometry optimized with the B3LYP/MWB model chemistry, a tetrahe-
dron, b triangular prism, c deformed pentagonal pyramid, d capped tetrahedron
The relative energy (a.u.) calculated with the ROHF/LANL2DZ method for Rh8
clusters are shown in Table 2. We searched the spin multiplicity up to 21. The
relative energy of isomers with different geometry is 23.76, 1.023, and 1.249 eV for
the cubic, bctp, and incomplete-ico respectively. In Fig. 2, we show the optimized
structures of the Rh8 clusters. Figure 2a is the lowest energy isomer, a bicapped
octahedron, bcoh, with a multiplicity equal to 19. Figure 2b corresponds to the
bicapped triangular prism isomer, bctp, with a multiplicity of 19. Figure 2c shows
an isomer with the incomplete icosahedron structure. Finally, a cubic isomer with
M = 13 is presented in Fig. 2d. The octahedron isomer with the minimum energy is
in complete agreement with the experimental results reported by D. J. Harding et al.
[10, 11], and by M. R. Beltran et al. [48], and with theoretical results obtained with
different methods such as local spin density [24, 57], molecular dynamics [26],
effective core potential [57], B3LYP [48, 49], which unlike of the reported here
Small Rhodium Clusters: A HF and DFT Study–III 217
Table 1 Rh6 relative energy (eV) obtained from B3LYP/MWB calculations for isomers of Rh6 in
different states of spin multiplicity, in bold the octahedron isomer
Multiplicity Triangular prism Pentagonal pyramid Square bi-pyramid
1 0.346 2.503 1.309
3 0.354 1.037 0.786
5 0.275 0.786 0.522
7 0.139 0.340 0.375
9 0 0.340 0
11 0.272 0 0.250
13 0.381 0.609 0.675
15 1.167 1.162 1.178
17 3.570 3.763 3.891
19 6.620 nc 7.132
21 nc nc nc
23 14.528 14.749 nc
nc not converged
were performed by keeping the symmetry, i.e. only optimized bond distances
[19, 56].
For the cluster formed by 13 rhodium atoms, taking as the initial geometry that
presented in Fig. 3a, we performed two series of calculations using density func-
tional theory. The first one using the Becke-Lee-Yang-Parr [58, 59] functional and a
double-Z basis set with effective core potential to represent the electrons close to the
nucleus, and the second using the PBE functional [44, 45]. Both sets of calculations
were carried out by varying the multiplicity from 2 to 26, performed with full
optimization, without symmetry constraints, and following the Berny algorithm for
minimization as it is implemented in the Gaussian 03 computer package.
218 M. A. Mora and M. A. Mora-Ramírez
Fig. 2 Isomers of Rh8 geometry in its minimum energy conformation obtained with the ROHF/
LANL2DZ. a bi-capped octahedron, b bi-capped triangular prism, c incomplete icosahedron,
d cube
Fig. 3 Rh13 with the functional B3LYP a configuration initial b optimized with M = 2 c the
minimum energy configuration for M = 22
been reported as optimized [27, 55, 57], and are markedly different from those
obtained by us using the same method of calculation. It seems that in those cal-
culations, geometries reported as the minimum energy were not optimized at the
DFT level, probably assuming predefined structures obtained by molecular dynamic
or similar methods.
The series of PBE calculations was also initiated with the geometry presented in
Fig. 3a, without any restrictions during the optimization process. The final geom-
etry, Fig. 4, is similar to that found with the B3LYP functional. The difference is
that the plane formed by the atoms 3, 4, 7, 8, 11, 12 is now more deformed, and the
cube formed by the atoms 1–8 is markedly deformed. Again, the electronic state of
minimum energy corresponds to M = 22.
This same type of calculation was performed earlier [22]. In that report, it is
mentioned that the minimum energy geometry is the so-called non-icosahedral one.
Since our initial geometry is not exactly that considered in those studies, we take as
the initial conformation a double single cube geometry, DSC, previously reported
[19, 28] as the minimum energy conformer. This initial conformation was used to
Small Rhodium Clusters: A HF and DFT Study–III 221
Fig. 6 Final geometry obtained for Rh13 with full optimization. The geometry obtained for both
PBE and B3LYP functionals are very similar
222 M. A. Mora and M. A. Mora-Ramírez
perform two other sets of calculations with the same B3LYP and PBE functionals.
Figure 5 shows the initial geometry, the final geometry obtained without any kind
of restriction for both functionals. Figure 6 is very similar. It is a simple double
distorted cube formed by oblique parallelepipeds. Although the geometries are very
similar, that obtained with the functional PBE shows shorter distances than the
corresponding one calculated with the B3LYP functional. The angle formed by the
distorted cubes is 140° and 124° for B3LYP and PBE respectively. The initial
geometry is not preserved during the optimization process. The geometry optimized
by us is 0.13 eV more stable with PBE, and 0.58 eV more stable with B3LYP, than
the capped double cube conformation previously reported. The energies are not the
only differences. Reference [56] reported a magnetic moment of 9 μB for the double
simple cubic, DSC, geometry with the PBE method while we obtain a magnetic
moment of 13 μB with the same method (PBE), and 21 μB using B3LYP, equal to
that obtained by Reddy et al. [18] using the von Barth-Hedin form of the
exchange-correlation contributions in the discrete variational method. It is
Fig. 7 Rh13 cluster in its minimum energy conformation, a centered icosahedron, with
multiplicity 16, obtained with the ROHF method
Small Rhodium Clusters: A HF and DFT Study–III 223
important to emphasize that the DSC geometry reported as the most stable is not
retained once it is subjected to a process of optimization without any kind of
restrictions.
Rh13 is one of the most studied [18–29, 38, 51, 60] clusters because it is
considered as the seed for different cluster growth patterns. A wide dispersion in the
calculated multiplicities is present in the literature.
The ROHF calculations for Rh13 converge to an interesting structure presented
in Fig. 7. It is an interlocking series of pentagons, an irregular icosahedron, with a
rhodium atom in the center. The icosahedron is composed of interlocking pen-
tagonal “caps”. Every vertex of the icosahedron is the top of a pentagonal cap. We
performed calculations for multiplicities in the range from 4 to 30. Table 4 shows
that the energy has an oscillating behavior. Three minimum energy states are
present at multiplicities of 10, 16 and 28. There is a small energy difference of 6.17
and 2.17 eV in relation to the multiplicity equal to 28.
In Fig. 7, the pentagons formed by the Rh1-Rh5, and the Rh6-Rh10 atoms have
different bond distances. These are 2.8 Å and 2.66 Å in average respectively. The
average distance from the vertex, Rh11 to the Rh1-Rh5 pentagon is 2.89 Å and the
distance from the vertex Rh13 to the pentagon Rh6-Rh10 is 3.8 Å. Also shown in
Table 4 is the dipole moment of the cluster in different electronic states.
There are two possibilities for the formation of the anion or the cation. The
cation could be formed when the neutral cluster loses an electron. It may be a
previously paired electron in which case the multiplicity of the cation is increased
by one unit. If the loss is that of a previously unpaired electron, the multiplicity
decreases by one unit in relation to the neutral cluster. On the other hand, the
Table 4 Rh13, energy (a.u.), relative energy (eV), dipolar moment (D), energy of the HOMO,
energy of the LUMO, and the band gap values obtained from ROHF calculations for different
states of spin multiplicity
Mult. Energy R. Energy D HOMO LUMO GAP
4 −1411.243 11.97 2.712 −0.031 −0.010 0.021
6 −1411.314 10.04 2.143 −0.065 −0.003 0.062
8 −1411.317 9.982 1.460 −0.051 −0.014 0.037
10 −1411.457 6.177 1.409 −0.035 −0.010 0.026
12 −1411.416 7.265 2.021 −0.037 0.002 0.039
14 −1411.423 7.075 1.978 −0.032 −0.004 0.028
16 −1411.604 2.176 0.995 −0.044 −0.002 0.042
18 −1411.588 2.585 1.132 −0.045 −0.001 0.041
20 −1411.579 2.830 2.872 −0.024 −0.004 0.020
22 −1411.618 1.769 1.082 −0.031 0.000 0.031
24 −1411.624 1.633 2.618 −0.025 −0.002 0.023
26 −1411.672 0.299 0.968 −0.026 −0.011 0.015
28 −1411.684 0.000 2.254 −0.036 −0.009 0.025
30 −1411.646 1.034 3.048 −0.033 −0.015 0.019
224 M. A. Mora and M. A. Mora-Ramírez
Table 5 Rh13 selected parameters ROHF/LANL2DZ for RH13 neutral with multiplicity 16 and
its ionic species
Specie Charge, Energy (a. E. A. I. P. HOMO LUMO GAP Dip. Mom.
multiplicity u.)
Rh13 +1, 15 −1411.399 0.206 −0.127 −0.115 0.012 0.85
+1, 17 −1411.516 0.087 −0.135 −0.120 0.016 1.05
0, 16 −1411.604 −0.444 −0.003 0.014 0.99
−1, 17 −1411.615 0.012 0.052 0.102 0.049 0.84
−1, 15 −1411.596 0.007 0.599 0.098 0.500 0.62
+1, 27 NC
+1, 29 NC
0, 28 −1411.684 −0.036 −0.009 0.027 2.254
−1, 27 −1411.654 −0.03 −0.046 0.096 0.05 2.656
−1, 29 −1411.641 −0.04 −0.036 −0.009 0.027 2.005
formation of the anion occurs through the gain of an electron. The new electron
may be unpaired in which case the multiplicity increases, or it might pair with a
previously unpaired electron in the neutral cluster thus decreasing the multiplicity.
Table 5 reports the Restricted Open shell Hartree Fock energies of
Rh13 M = 16 and M = 28, and its ionic clusters, the charge and multiplicity, the
electron affinities (EA) and the ionization potential (IP). The electron affinities were
calculated as the energy difference between the neutral and the anionic clusters
while the ionization potential was calculated as the energy difference between the
cation and the neutral molecule. On the other hand the Ionization Potential calcu-
lated from the HOMO following Koopman’s theorem, is 0.116 a.u. = 3.15 eV for
M = 16.
4 Conclusions
We report a study of rhodium clusters using ROHF and DFT methods with full
optimization.
Our results agree with those of other researchers concerning the equilibrium
geometry, but disagree on the magnetic ground state
• The equilibrium structure of the isomers is unambiguously determined with
ROHF methods. For Rh6 and Rh8, the ROHF minimum energy conformation is
in excellent agreement with experiment.
• Different spin states are quite close in energy. All of them have essentially the
same equilibrium structure.
• With different XC functionals, different spin states are obtained for the same
conformation.
Small Rhodium Clusters: A HF and DFT Study–III 225
References
23. Sun Y, Zhang M, Fournier R (2008) Periodic trends in the geometric structures of 13-atoms
metal clusters. Phys Rev B 77:0754435
24. Jinlong Y, Toigo F, Kelin W (1994) Structural electronic, and magnetic properties of small
rhodium clusters. Phys Rev B 50:7915–7924
25. Reddy BV, Nayak SK, Khanna SN, Rao BK, Jena P (1999) Electronic structure and
magnetism of Rhn (n = 2-13) clusters. Phys Rev B 59:5214–5222
26. Guirado-López R, Villaseñor-González P, Dorantes-Dávila J, Pastor GM (2000) Magnetism
of Rhn clusters. J Appl Phys 87:4906
27. Aguilera-Granja F, Montejano-Carrizales JM, Guirado-López RA (2006) Magnetic properties
of small 3d and 4d transition metal clusters: the role of a noncompact growth. Phys Rev B
73:115422
28. Bae YC, Osansi H, Kumar V, Kawazone Y (2004) Nonicosahedral growth and magnetism
behavior of rhodium clusters. Phys Rev B 70:195413–195419
29. Rogan J, García G, Loyola C, Orellana W, Ramírez R, Kiwi M (2006) Alternative search
strategy for minimal energy nanocluster structures: the case of rhodium, palladium, and silver.
J Chem Phys 125:214708–214712
30. Joswig JO, Springborg M (2003) Generic algorithms search for global minima of aluminium
clusters using Sutton-Chen potential. Phys Rev B 68:085408
31. Zhan L, Chen JZY, Montejano-Carrizalez JM, Guirado-López RA (2006) Phys Rev B
73:115422
32. Aprá E, Ferrando R, Fontunelli A (2006) Density-functional global optimization of gold
clusters. Phys Rev B 73:205414
33. Wales DJ, Doye JPK (1997) Global optimization by basin-hopping and the lowest energy
structures of Lennard-Jones clusters containing up to 110 atoms. J Phys Chem A 101:
5111–5116
34. Kim HG, Choi SK, Lee HM (2008) New algorithm in the basin-hopping Monte Carlo to find
the global minimum structure of unary and binary metallic nano clusters. J Chem Phys
128:144702
35. Erkoc S, Saltaf R (1999) Monte Carlo computer simulations of copper clusters. Phys Rev A
60:3053
36. Cheng J, Fournier R (2004) Structural optimization of atomic clusters by Tabu search in
descriptor spaces. Theor Chem Acc 112:7–15
37. Baletto F, Ferrando R (2005) Structural properties of nanoclusters: energetic, thermodynamic,
and kinetic effects. Rev Mod Phys 77:371
38. Piotrowsky MJ, Piquini P, Da Silva JLF (2010) Density functional theory investigation of 3d,
4d, and 5d 13-atom metal clusters. Phys Rev B 81:155446–155459
39. Piotrowski MJ, Piquini P, Odashima MM, DaSilva JLF (2011) Transition-metal 13-atom
clusterts assessed with solid and surface-biased functionals. J Chem Phys 134:134105–
134110
40. Hay PJ, Wadt WR (1985) Ab initio effective core potentials for molecular calculations.
Potentials for main group elements Na to Bi. J Chem Phys 82:270
41. Wadt WR, Hay PJ (1985) Ab initio effective core potentials for molecular calculations.
Potentials for the transition metal atoms Sc to Hg. J Chem Phys 82:284
42. Hay PJ, Wadt WR (1985) Ab initio effective core potentials for molecular calculations.
Potentials for K to Au including the outermost core orbitals. J Chem Phys 82:299
43. Wood JH, Boring AM (1978) Improved Pauli Hamiltonian for local-potential problems. Phys
Rev B 18:2701
44. Perdew JP, Burke K, Ernzerhof M (1996) General gradient approximation made simple. Phys
Rev Lett 77:3865
45. Perdew JP, Burke K, Ernzerhof M (1997) General gradient approximation made simple [Phys
Rev Lett 77:4884(1996)] Phys Rev Lett 78:1396
46. Peng C, Ayala PY, Schelegel HB, Frish MJ (1996) Using redundant internal coordinates to
optimize equilibrium geometries and transition states. J Comput Chem 17:49
Small Rhodium Clusters: A HF and DFT Study–III 227
47. Peng C, Schelegel HB (1994) Combining synchronous transit and quasi-Newton methods to
find transit states. Israel J Chem 33:449
48. Beltrán MR, Buendía Zamudío F, Chanhan V, Sen P, Wang H, Ko YJ, Bowen K (2013) Ab
initio and anion photoelectron studies of Rhn (n = 1-9) clusters. Eur Phys J D 67:63–70
49. Bertin V, Lopez-Rendón R, del Angel G, Poulain E, Avilés R, Uc-Rosas V (2010)
Comparative theoretical study of small Rhn nanoparticles (2 ≤ n ≤ 8) using DFT methods.
Int J Quantum Chem 110:1152–1164
50. Harding DJ, Davies RDL, Mackenzie SR, Walsh TR (2008) Oxides of small rhodium clusters:
theoretical investigation of experimental reactivities. J Chem Phys 129:124304–124310
51. Mora MA, Mora-Ramírez MA, Rubio-Arroyo Manuel F (2010) Structural and electronic
study of neutral, positive and negative small rhodium clusters [Rhn, Rh+ -
n , Rhn]. Int J Quantum
Chem 110:2541–2547
52. Li ZQ, Yu JZ, Ohno K, Kawazoe Y (1999) Calculations on the magnetic properties of
rhodium clusters. J Phys Condens Matter 7:47–53
53. Hang TD, Hung HM, Thiem LN, Nguyen HMT (2015) Electronic structure and
thermochemical properties of neutral and anionic rhodium clusters Rhn, n = 2-13. Evolution
of structures and stabilities of binary clusters RhmM (M = Fe Co, Ni; m = 1-6). Comput
Theor Chem 1068:30–41
54. Chien CH, Blaisten-Barojas E, Pedersen MR (1998) Magnetic and electronic properties of
rhodium clusters. Phys Rev A 58:2196–2202
55. Harding D, Mackenzie SR, Walsh TR (2006) Structural isomers and reactivity for Rh6, Rh+ 6.
J Phys Chem B 110:18272–18277
56. Da Silva JLF, Piotrowski MJ, Aguilera-Granja F (2012) Phys Rev B 86:125430–125435
57. Sun Y, Fournier R, Zhang M (2009) Structural and electronic properties of 13-atom 4d
transition-metal clusters. Phys Rev B 79:043202–043211
58. Lee C, Yang W, Parr RG (1988) Development Colle-Salvetti correlation-energy formula into
a functional of the electron density. Phys Rev B 37:785
59. Miehlich B, Sabin A, Stoll H, Preus H (1989) Results obtained with the correlation energy
density functional of Becke and Lee, Yang and Parr. Chem Phys Lett 157:200–206
60. Futschek T, Marsman M, Hafner J (2005) Structural and magnetic isomers of small Pd and Rh
clusters: an ab initio functional study. J Phys Condens Matter 17:5927–5963
Spectroscopy of Radiative Decay
Processes in Heavy Rydberg
Alkali Atomic Systems
Abstract We present the results of studying the radiation decay processes and
computing the probabilities and oscillator strengths of radiative transitions in spectra
of heavy Rydberg alkali-metal atoms. All calculations of the radiative decay
(transitions) probabilities have been carried out within the generalized relativistic
energy approach (which is based on the Gell-Mann and Low S-matrix formalism)
and the relativistic many-body perturbation theory with using the optimized
one-quasiparticle representation and an accurate accounting for the critically
important exchange-correlation effects as the perturbation theory second and higher
orders ones. The precise data on spectroscopic parameters (energies, reduced dipole
transition matrix elements, amplitude transitions) of the radiative transitions nS1/2
→ n′P1/2,3/2 (n = 5, 6; n′ = 10–70), nP1/2,3.2 → n′D3/2,5/2 (n = 5, 6; n′ = 10–80) in
the Rydberg Rb, Cs spectra and the transitions 7S1/2-nP1/2,3/2, 7P1/2,3.2-nD3/2,5/2
(n = 20–80) in the Rydberg francium spectrum are presented. The obtained results
are analyzed and discussed from viewpoint of the correct accounting for the rela-
tivistic and exchange-correlation effects. It has been shown that theoretical approach
used provides an effective accounting of the multielectron exchange-correlation
effects, including effect of essentially non-Coulomb grouping of Rydberg levels and
others.
1 Introduction
Rydberg francium spectrum are listed. The data are discussed from the viewpoint of
the correct accounting for the relativistic and exchange-correlation effects. It has
been shown that the theoretical approach used provides a precise accounting for the
important exchange-correlation effects, including the effect of essentially
non-Coulomb grouping of Rydberg levels, continuum pressure etc.
All calculations of the radiative decay (transitions) probabilities (matrix ele-
ments) in the studied atomic systems have been performed with using the gener-
alized relativistic energy approach and the relativistic many-body perturbation
theory (PT) with using the optimized one-quasiparticle representation and an
accurate accounting of the exchange-correlation effects, including the effect of
essentially non-Coulomb grouping of Rydberg levels [61–63].
Let us remind that the theoretical fundamentals of an energy approach in a case of
the one-electron ions have been considered by Labzovsky et al. [57, 58]. Originally
the energy approach to radiative and autoionization processes in multielectron atoms
and ions has been developed by Ivanova-Ivanov et al. [59–62, 64–67]. More
accurate, advanced version of the relativistic energy approach has been further
developed in Refs. [63, 68–72]. The energy approach is based on the Gell-Mann and
Low S-matrix formalism combined with the relativistic perturbation theory. In
relativistic case the Gell-Mann and Low formula expressed an energy shift ΔE
through the electrodynamical scattering matrix including interaction with as the
photon vacuum field as a laser field. The first case is corresponding to determination
of radiative decay characteristics for atomic systems. Earlier we have applied the
corresponding generalized versions of the energy approach to many problems of
atomic, nuclear and even molecular spectroscopy, including, cooperative
electron-gamma-nuclear “shake-up” processes, electron-muon-beta-gamma-nuclear
spectroscopy, spectroscopy of atoms in a laser field etc. [73–98].
Let us describe in brief the key moments of our theoretical approach (look for more
details in Refs. [63, 65–69, 73–76]). As usually, the wave functions zeroth basis is
found from the Dirac equation solution with self-consistent total potential.
The bare Hamiltonian is as follows:
1 − αi αj
H = ∑ fαcp − βmc2 + Uðri jZÞg + + ∑ exp iωij rij ⋅ , ð1Þ
i i>j rij
where αi , αj —the Dirac matrices, ωij —the transition frequency, c—the light
velocity, Z is a charge of the atomic nucleus. Within relativistic perturbation theory
[3, 4] we introduce the zeroth–order Hamiltonian as:
232 V. B. Ternovsky et al.
H0 = ∑ fαcpi − βmc2 + ½ − Z ̸ri + UMF ðri jbÞ + VXC ðri Þg, ð2Þ
i
ΔE = ReΔE + iΓ ̸ 2 ð3Þ
e2 jωαn j
ImΔEðBÞ = − ∑ Vαnαn , ð4Þ
4π
α>n>f
α>n≤f
where (α > n > f) for electron and (α < n < f) for vacancy. The matrix element is
determined as follows:
ZZ
jωj sinjωjr12
Vijkl = dr1 dr2 Ψ *i ðr1 ÞΨ *j ðr2 Þ ð1 − α1 α2 ÞΨ *k ðr2 ÞΨ *l ðr1 Þ ð5Þ
r12
Spectroscopy of Radiative Decay Processes … 233
When calculating the matrix elements (5), one should use the angle symmetry of
the task and write the corresponding expansion for sinjωjr12 ̸ r12 on spherical
harmonics as follows [62]:
sin jωjr12 π ∞
= pffiffiffiffiffiffiffiffi ∑ ðλÞJλ + 1 ̸2 ðjωjr1 ÞJλ + 1 ̸2 ðjωjr2 ÞPλ ðcos rd
1 r2 Þ ð6Þ
r12 2 r1 r2 λ = 0
where ji is the total single electron momentums, mi—the projections; QQul is the
Coulomb part of interaction, QBr—the Breit part.
The total radiation width of the one-quasiparticle state can be presented in the
following form:
ΓðγÞ = − 2 ImM 1 ðγÞ = − 2 ∑ ð2j + 1ÞImQλ nγ lγ jγ nlj
λnlj
Qλ = QCul Br ð8Þ
λ + Qλ .
QBr Br Br
λ = Qλ, λ − 1 + Qλ, λ + QBr
λ, λ + 1
The individual terms of the Σnlj sum correspond to the partial contribution of the
nλ lλ jλ → nlj transitions; Σλ is a sum of the contributions of the different multiplicity
transitions. The detailed expressions for the Coulomb and Breit parts can be found
in Refs. [62–66].
The imaginary parts of the Coulomb part QCul λ and the Breit part contain the
radial Rλ and angular Sλ integrals as follows (in the Coulomb units) [65]:
n
−1 e e e e
Im QCul
λ ð 12; 43 Þ = Z Im Rλ ð 12; 43 ÞS λ ð 12; 43 Þ + Rλ 12; 4 3 S λ 12; 4 3 +
o ð9Þ
+ Rλ 1e 2; e43 Sλ 1e 43 + Rλ e
2; e 1e2; e
4e3 Sλ e 1e
2; e
4e3 .
1 n e e
l e e
ImQBr
λ, l = Im Rλ 12; 4 3 Sλ 12; 4 3 + Rλ e1e2; 43 Slλ e1e2; 43 +
Z
o ð10Þ
+ Rλ e
12; e
43 Sl e
12; e
λ 43 + Rλ e1e
2; e
4e3 Sl e1e2; e
4e3 .
λ
Here fλ l1 l3 g means that λ, l1 and l3 must satisfy the triangle rule and the sum
λ + l1 + l3 must be an even number. The rest terms in (9), (10) include the small
components of the Dirac functions. The tilde designates that the large radial
234 V. B. Ternovsky et al.
component f must be replaced by the small one g, and instead of li , lĩ = li − 1 should
be taken for ji < li and lĩ = li + 1 for ji > li . The detailed definitions for the radial Rλ
and angular Sλ integrals can be found in Refs. [59, 64–67].
The total probability of a λ—pole transition is usually represented as a sum of
the electric PEλ and magnetic PM λ parts. The electric (or magnetic) λ—pole transition
γ → δ connects two states with parities which by λ (or λ +1) units. In our desig-
nations (the radiative γ → δ transition) one could write:
ð12aÞ
1 ̸ 3 Z
1 ̸ 3
⟨ ρð0Þ
c ⟩= dr ρð0Þ
c ðrÞ θðrÞ, ð12bÞ
h i2 1 ̸2
̸3
θðrÞ = 1 + 3π 2 ⋅ ρð0Þ
c ðrÞ ̸c 2
ð12cÞ
Spectroscopy of Radiative Decay Processes … 235
where ρ0c is the core electron density (without account for the quasiparticle), X is
numerical coefficient, c is the light velocity. The contribution of the ladder diagrams
(these diagrams describe the immediate interparticle interaction) is summarized by a
modification of the perturbation theory zeroth approximation mean-field central
potential (look [35, 65]), which include the screening (anti-screening) of the core
potential of each particle by the two others. The details of calculating this contri-
bution can be found in Refs. [63, 65–76]. All computing was performed with using
the modified PC code “Superatom-ISAN” (version 93).
In Tables 1 and 2 we present the experimental and theoretical values (in atomic
units: a.u.) of the reduced dipole transition matrix elements for the Fr and Cs atoms:
experimental data—Exp; theoretical data: perturbation theory (PT)-DFSD—PT with
the Dirac-Hartree-Fock zeroth approximation (single-double SD approximation in
which single and double excitations of Dirac-Hartree-Fock wave functions are
Table 1 The reduced dipole transition matrix elements for Fr (see text)
Transition/ PT-DFSD PT-DFSD EMP PT-RHF PT-RHF DF RPT-EA Exp.
(corr)
method (corr)
7p1/2-7s 4.256 – – 4.279 4.304 4.179 4.272, 4.277
4.274
8p1/2-7s 0.327 0.306 0.304 0.291 0.301 – 0.339
9p1/2-7s 0.110 0.098 0.096 – – – 0.092
10p1/2-7s – – – – – – 0.063
7p3/2-7s 5.851 – – 5.894 5.927 5.791 5.891 5.898
8p3/2-7s 0.934 0.909 0.908 0.924 – – 0.918 –
9p3/2-7s 0.436 0.422 0.420 – – – 0.426 –
10p3/2-7s – – – – – – 0.284 –
7p1/2-8s 4.184 4.237 4.230 4.165 4.219 4.196 4.228 –
8p1/2-8s 10.02 10.10 10.06 10.16 10.00 10.12 –
9p1/2-8s 0.985 – 0.977 – – – 0.972 –
10p1/2-8s – – – – – – 0.395 –
7p3/2-8s 7.418 7.461 7.449 7.384 7.470 7.472 7.453 –
8p3/2-8s 13.23 13.37 13.32 13.45 13.26 13.35 –
9p3/2-8s 2.245 – 2.236 – – – 2.232 –
10p3/2-8s – – – – – – 1.058 –
7p1/2-9s 1.016 – 1.010 – – – 1.062 –
8p1/2-9s 9.280 – 9.342 – – – 9.318 –
9p1/2-9s 17.39 – 17.40 – – – 17.42 –
(continued)
236 V. B. Ternovsky et al.
Table 1 (continued)
Transition/ PT-DFSD PT-DFSD EMP PT-RHF PT-RHF DF RPT-EA Exp.
(corr)
method (corr)
10p1/2-9s – – – – – 1.836 –
7p3/2-9s 1.393 – 1.380 – – – 1.41 –
8p3/2-9s 15.88 – 15.92 – – – 15.96 –
9p3/2-9s 22.59 – 22.73 – – – 22.68 –
10p3/2-9s – – – – – – 3.884 –
Table 2 The reduced dipole transition matrix elements for Cs (see text)
Transition PT-DFSD PT-DFSD DF PT-RHF QDA EF-RMP Exp.
(corr)
6p1/2-6s 4.482 4.535 4.510 – 4.282 4.489 4.4890(7)
6p3/2-6s 6.304 6.382 6.347 – 5.936 6.323 6.3238(7)
7p1/2-6s 0.297 0.279 0.280 0.2825 0.272 0.283 0.284(2)
7p3/2-6s 0.601 0.576 0.576 0.582 0.557 0.583 0.583(9)
8p1/2-6s 0.091 0.081 0.078 – 0.077 0.088 –
8p1/2-6s 0.232 0.218 0.214 – 0.212 0.228 –
6p1/2-7s 4.196 4.243 4.236 4.237 4.062 4.234 4.233(22)
6p3/2-7s 6.425 6.479 6.470 6.472 6.219 6.480 6.479(31)
7p1/2-7s 10.254 10.310 10.289 10.285 9.906 10.309 10.309
(15)
7p3/2-7s 14.238 14.323 14.293 14.286 13.675 14.323 14.325
(20)
Fig. 1 A dependence of the calculated reduced dipole matrix elements upon principal quantum
number for Rydberg atom Rb: 5P3/2 → nD5/2 (n ∼ 70). The available experimental data are listed
as a circle; Theory: continuous line—our data, dotted line- data by Piotrowicz et al. within the
quasi-classical Dyachkov-Pankratov model (see text)
Fig. 2 A dependence of the calculated reduced dipole matrix elements upon principal quantum
number for Rydberg atom Cs: 6P3/2 → nD5/2 (n ∼ 70). The available experimental data are listed
as a circle; Theory: continuous line—our data, dotted line- data by Piotrowicz et al. within the
quasi-classical Dyachkov-Pankratov model (see text)
Fig. 3 A dependence of the calculated reduced dipole matrix elements upon principal quantum
number for Rydberg atom Fr: 7P3/2 → nD5/2, n = 10–80 (our data)
Acknowledgements The authors are very much thankful to Prof. J. Maruani and Dr. Y. A. Wang
for invitation to make contributions on the QSCP-XVI workshop (Vancouver, Canada). The useful
comments of the anonymous referees are very much acknowledged too.
References
1. Martin W (2004) NIST spectra database, version 2.0. NIST, Washington. http://physics.nist.
gov.asd
2. Moore C (1987) NBS spectra database. NBS, Washington
3. Weiss A (1977) J Quant Spectrosc Radiat Transf 18:481; Phys Scripta T65:188 (1993)
4. Nadeem A, Haq SU (2010) Phys Rev A 81:063432; (2011) Phys Rev A 83:063404
5. Simsarian JE, Orozco LA, Sprouse GD, Zhao WZ (1998) Phys Rev A 57:2448
6. Curtis L (1995) Phys Rev A 51:4574
7. Li Y, Pretzler G, Fill EE (1995) Phys Rev A 52:R3433–R3435
8. Feng Z-G, Zhang L-J, Zhao J-M, Liand C-Y, Jia S-T (2009) J Phys B At Mol Opt Phys
42:145303
9. Marinescu M, Vrinceanu D, Sadeghpour HR (1998) Phys Rev A 58:R4259
10. Safronova UI, Johnson W, Derevianko A (1999) Phys Rev A 60:4476; Dzuba V,
Flambaum V, Safranova MS (2006) Phys Rev A 73:02211
11. Grant IP (2007) Relativistic quantum theory of atoms and molecules, theory and
computation. In: Springer series on atomic, optical, and plasma physics, vol 40. Springer,
Berlin, pp 587–626
12. Glushkov AV (2008) Relativistic quantum theory. Quantum mechanics of atomic systems.
Astroprint, Odessa, p 700
Spectroscopy of Radiative Decay Processes … 239
13. Wilson S (2007) Recent advances in theoretical physics and chemistry systems. In:
Maruani J, Lahmar S, Wilson S, Delgado-Barrio G (eds) Series: Progress in theoretical
chemistry and physics, vol 16. Springer, Berlin, pp 11–80
14. Feller D, Davidson ER (1981) J Chem Phys 74:3977
15. Froelich P, Davidson ER, Brändas E (1983) Phys Rev A 28:2641
16. Rittby M, Elander N, Brändas E (1983) Int J Quantum Chem 23:865
17. Wang YA, Yam CY, Chen YK, Chen GH (2011) J Chem Phys 134:241103
18. Maruani J (2016) J Chin Chem Soc 63:33
19. Pavlov R, Mihailov L, Velchev C, Dimitrova-Ivanovich M, Stoyanov Z, Chamel N, Maru-
ani J (2010) J Phys Conf Ser 253:012075
20. Dietz K, Heβ BA (1989) Phys Scripta 39:682
21. Kohn JW, Sham LJ (1964) Phys Rev A 140:1133
22. Gross EG, Kohn W (2005) Exchange-correlation functionals in density functional theory.
Plenum, New York
23. Froese Fischer C, Tachiev G, Irimia A (2006) GAtom Data Nucl Data Tables 92:607
24. Cheng K, Kim Y, Desclaux J (1979) At Data Nucl Data Tables 24:11
25. Safranova UI, Safranova MS, Johnson W (2005) Phys Rev A 71:052506
26. Ternovsky EV, Antoshkina OA, Florko TA, Tkach TB (2017) Photoelectronics 26:139
27. Indelicato P, Desclaux JP (1993) Phys Scripta T 46:110
28. Dzuba VA, Flambaum VV, Sushkov OP (1995) Phys Rev A 51:3454
29. Sapirstein J (1998) Rev Modern Phys 70:55
30. Piotrowicz M, MacCormick C, Kowalczyk A et al (2011) arXiv:1103.0109v2 [quant-ph];
Beterov I, Mansell CW, Yakshina EA et al (2012) arXiv:1207.3626v1 [physics.atom-ph]
31. Dyachkov LG, Pankratov PM (1994) J Phys B 27(3):461; Piotrowicz M, MacCormick C,
Kowalczyk A, Bergamini S, Yakshina EA (2011) New Journ Phys 13:093012
32. Quinet P, Argante C, Fivet V, Terranova1 C, Yushchenko AV, Biémont É (2007) Astrophys
Astron 474:307
33. Biémont É, Fivet V, Quinet P (2004) J Phys B At Mol Opt Phys 37:4193
34. Glushkov AV (1991) Opt Spectrosc 70:555
35. Glushkov AV (1990) Soviet Phys J 33(1):1
36. Khetselius OY (2009) Int J Quantum Chem 109:3330
37. Khetselius OY (2009) Phys Scripta T135:014023
38. Malinovskaya SV, Glushkov AV, Khetselius OY, Svinarenko AA, Mischenko EV,
Florko TA (2009) Int J Quant Chem 109(4):3325
39. Glushkov AV, Loboda AV, Gurnitskaya EP, Svinarenko AA (2009) Phys Scripta
T135:014022
40. Glushkov AV, Khetselius OY, Svinarenko AA (2013) Phys Scripta T153:014029
41. Svinarenko AA (2014) J Phys Conf Ser 548:012039
42. Glushkov AV, Mansarliysky VF, Khetselius OY, Ignatenko AV, Smirnov A, Prepelitsa GP
(2017) J Phys Conf Ser 810:012034
43. Glushkov AV, Loboda AV (2007) J Appl Spectrosc 74:305. (Springer)
44. Malinovskaya SV, Glushkov AV, Khetselius OY, Loboda AV, Lopatkin YuM,
Svinarenko AA, Nikola LV, Perelygina TB (2011) Int J Quantum Chem 111:288
45. Glushkov AV, Khetselius OY, Loboda AV, Ignatenko AV, Svinarenko AA,
Korchevsky DA, Lovett L (2008) Spectral line shapes. AIP Conf Proc 1058:175
46. Khetselius OY (2012) J Phys Conf Ser 397:012012
47. Khetselius OY, Florko TA, Svinarenko AA, Tkach TB (2013) Phys Scripta T153:014037
48. Khetselius OY (2008) Spectral line shapes. AIP Conf Proc 1058:363
49. Khetselius OY, Glushkov AV, Gurnitskaya EP, Loboda AV, Mischenko EV, Florko T,
Sukharev D (2008) Spectral line shapes. AIP Conf Proc 1058:231
50. Khetselius OY (2007) Photoelectronics 16:129
51. Khetselius OY, Gurnitskaya EP (2006) Sens Electron Microsyst Technol N 3:35
52. Khetselius OY, Gurnitskaya EP (2006) Sens Electron Microsyst Technol N 2:25
53. Johnson WR, Lin CD, Cheng KT (1980) Phys Scr 21:409
240 V. B. Ternovsky et al.
86. Glushkov AV, Khetselius OY, Lovett L (2010) Advances in the theory of atomic and
molecular systems dynamics, spectroscopy, clusters, and nanostructures. In: Piecuch P,
Maruani J, Delgado-Barrio G, Wilson S (eds) Series: Progress in theoretical chemistry and
physics, vol 20. Springer, pp 125–152
87. Glushkov AV, Khetselius OY, Loboda AV, Svinarenko AA (2008) Frontiers in quantum
systems in chemistry and physics. In: Wilson S, Grout PJ, Maruani J, Delgado-Barrio G,
Piecuch P (eds) Series: Progress in theoretical chemistry and physics, vol 18. Springer,
pp 543–560
88. Glushkov AV, Khetselius O, Gurnitskaya E, Loboda A, Florko T, Sukharev D, Lovett L
(2008) Frontiers in quantum systems in chemistry and physics. In: Wilson S, Grout P,
Maruani J, Delgado-Barrio G, Piecuch P (eds) Series: Progress in theoretical chemistry and
physics, vol 18. Springer, pp 507–524
89. Glushkov AV, Ambrosov SV, Loboda AV, Gurnitskaya EP, Khetselius OY (2006) Recent
advances in theoretical physics and chemistry systems. In: Julien JP, Maruani J, Mayou D,
Wilson S, Delgado-Barrio G (eds) Series: Progress in theoretical chemistry and physics, vol
15. Springer, pp 285–299
90. Malinovskaya SV, Glushkov AV, Dubrovskaya YV, Vitavetskaya LA (2006) Recent
advances in theoretical physics and chemistry systems. In: Julien J-P, Maruani J, Mayou D,
Wilson S, Delgado-Barrio G (eds) Series: Progress in theoretical chemistry and physics, vol
15. Springer, pp 301–307
91. Khetselius OY (2012) Quantum systems in chemistry and physics: progress in methods and
applications. In: Nishikawa K, Maruani J, Brandas E, Delgado-Barrio G, Piecuch P
(eds) Series: Progress in theoretical chemistry and physics, vol 26. Springer, pp 217–229
92. Khetselius OY (2015) Frontiers in quantum methods and applications in chemistry and
physics. In: Nascimento M, Maruani J, Brändas E, Delgado-Barrio G (eds) Series: Progress
in theoretical chemistry and physics, vol 29. Springer, pp 55–76
93. Glushkov AV, Svinarenko AA, Khetselius OY, Buyadzhi VV, Florko TA, Shakhman AN
(2015) Frontiers in quantum methods and applications in chemistry and physics. In:
Nascimento M, Maruani J, Brändas E, Delgado-Barrio G (eds) Series: Progress in theoretical
chemistry and physics, vol 29. Springer, pp 197–217
94. Khetselius OY, Zaichko PA, Smirnov AV, Buyadzhi VV, Ternovsky VB, Florko TA,
Mansarliysky VF (2017) Quantum systems in physics, chemistry, and biology. In: Tadjer A,
Pavlov R, Maruani J, Brändas E, Delgado-Barrio G (eds) Series: Progress in theoretical
chemistry and physics, vol 30. Springer, pp 271–281
95. Glushkov AV, Rusov VD, Ambrosov SV, Loboda AV (2003) New projects and new lines of
research in nuclear physics. In: Fazio G, Hanappe F (eds). World Scientific, Singapore,
pp 126–132
96. Glushkov AV, Malinovskaya SV, Loboda AV, Shpinareva IM, Prepelitsa GP (2006) J Phys
Conf Ser 35:420
97. Glushkov AV, Malinovskaya SV, Chernyakova YG, Svinarenko AA (2004) Int J Quantum
Chem 99:889
98. Glushkov AV, Ambrosov SV, Ignatenko AV, Korchevsky DA (2004) Int J Quantum Chem
99:936
99. Glushkov AV, Ignatenko AV, Khetselius OY, Ternovsky VB (2017) Spectroscopy of
Rydberg atoms and relativistic quantum chaos, OSENU, Odessa, p 152
100. Ternovsky VB, Buyadzhi VV, Gurskaya MY, Kuznetsova AA (2015) Preprint OSENU,
NAM-3 (Odessa, 2015), p 32; Preprint OSENU, OSENU, NAM-4 (Odessa, 2015), p 36
Enhancement Factors for Positron
Annihilation on Valence and Core
Orbitals of Noble-Gas Atoms
1 Introduction
Low-energy positrons annihilate in atoms and molecules forming two 𝛾 rays whose
Doppler-broadened spectrum is characteristic of the electron velocity distribution in
the states involved, and thus of the electron environment. This makes positrons a
unique probe in materials science. For example, vacancies and defects in semicon-
ductors and other industrially important materials can be studied [1–6]. Positron-
induced Auger-electron spectroscopy (PAES) [7–11] and time-resolved PAES
[11, 12] enable studies of surfaces with extremely high sensitivity, including dynam-
ics of catalysis, corrosion, and surface alloying [13]. The 𝛾 spectra are also sensitive
to the positron momentum at the instant of annihilation. This is exploited in Age-
1
Positron annihilation with core electrons is also affected by exchange-assisted tunnelling [26, 27].
This is a manifestation of electron exchange, which increases the wavefunctions of inner electrons in
the range of distances of the valence electrons. For this effect to be properly included in a calculation,
one needs to use true nonlocal exchange potentials, e.g., at the Hartree-Fock level, as is the case in
the present calculations.
Enhancement Factors for Positron Annihilation on Valence . . . 245
noble-gas atoms [19], with excellent agreement between the theoretical results and
experimental scattering cross sections and annihilation rates. The MBT work was
extended recently [24] to the 𝛾-spectra for (thermal) positron annihilation on noble-
gas atoms. It provided an accurate description of the measured spectra for Ar, Kr and
Xe [3] and firmly established the relative contributions of various atomic orbitals to
the spectra. The calculations also yielded “exact” ab initio EF 𝛾̄nl for individual elec-
tron orbitals nl, and found that they follow a simple scaling with the orbital ionization
energy [24].
In this work we provide a more detailed analysis and report EF for annihilation of
s-, p- and d-wave positrons with momenta up to the positronium formation threshold.
We demonstrate that the EF for a given electron orbital and positron partial wave are
insensitive to the positron momentum (in spite of the strong momentum dependence
of the annihilation probability [19]). Moreover, we show that whilst the EF for the
core orbitals are almost independent of the positron angular momenta, those for the
valence subshells vary between the positron s, p and d waves. In addition to their
use in correcting IPA calculations of positron annihilation with core electrons in
condensed matter, the positron-momentum dependent EF calculated here can be used
to determine accurate pick-off annihilation rates for positronium in noble gases [39].
2.1 Basics
Consider annihilation of a low-energy (𝜀 ∼ 1 eV) positron with momentum 𝐤 in a
many-electron system, e.g., an atom. In the dominant process, the positron anni-
hilates with an electron in state n to form two 𝛾-ray photons of total momentum
𝐏 [40]. In the centre-of-mass frame, where the total momentum 𝐏 is zero, the
two photons are emitted in opposite directions and have equal energies E𝛾 = p𝛾 c =
mc2 + 12 (Ei − Ef ) ≃ mc2 ≃ 511 keV, where Ei and Ef denote the energy of the initial
and final states of the system (excluding rest mass). When 𝐏 is non-zero, however,
the two photons no longer propagate in exactly opposite directions and their energy
is Doppler shifted. For example, for the first photon E𝛾1 = E𝛾 + mcV cos 𝜃, where 𝜃
is the angle between the momentum of the photon and the centre-of-mass velocity of
the electron-positron pair 𝐕 = 𝐏∕2m (assuming that V ≪ c, and p𝛾1 = E𝛾1 ∕c ≈ mc).
The Doppler shift of the photon energy from the centre of the line then is
Pc
𝜖 = E𝛾1 − E𝛾 = mc V cos 𝜃 = cos 𝜃. (1)
2
The typical momenta of electrons bound with√energy 𝜀n determine the characteristic
width of the annihilation spectrum 𝜖 ∼ Pc ∼ |𝜀n |mc2 ≫ |𝜀n |. Hence the shift 𝜀n ∕2
of the line centre E𝛾 from mc2 = 511 keV can usually be neglected, even for the core
electrons. The 𝛾 spectrum averaged over the direction of emitted photons (or that of
the positron momentum 𝐤) takes the form (see, e.g., [20])
246 D. G. Green and G. F. Gribakin
∞
1 PdPd𝛺𝐏
wn (𝜖) = |An𝐤 (𝐏)|2 , (2)
c ∫ ∫2|𝜖|∕c (2𝜋)3
where An𝐤 (𝐏) is the annihilation amplitude, whose calculation using MBT is
described below. The quantity |An𝐤 (𝐏)|2 is the annihilation momentum density.2
The annihilation rate 𝜆 for a positron in a gas of atoms or molecules with number
density nm is usually parametrized by
where r0 = e2 ∕mc2 is the classical radius of the electron (in CGS units) and Zef f
is the effective number of electrons per target atom or molecule that contribute to
∑
annihilation [42, 43]. It is found as a sum over electron states Zef f = n Zef f ,n , where
d3 𝐏
Zef f ,n = wn (𝜖) d𝜖 = |An𝐤 (𝐏)|2 (4)
∫ ∫ (2𝜋)3
is the partial contribution due to positron annihilation with electron in state n, and
where it is assumed that the incident positron wavefunction used in the calculation
of An𝐤 (𝐏) is normalized to a plane wave. In general, the parameter Zef f is different
from the number of electrons in the target atom Z. In particular, positron-atom and
electron-positron correlations can make Zef f ≫ Z [19, 24, 44–46].
where 𝛿𝓁 is the scattering phaseshift [47], Y𝓁m is the spherical harmonic, and
where the radial function with orbital angular momentum 𝓁 is normalized by
its asymptotic behaviour P𝜀𝓁 (r) ≃ (𝜋k)−1∕2 sin(kr − 𝜋𝓁∕2 + 𝛿𝓁 ). In the simplest
approximation the radial wavefunctions are calculated in the static field of the
ground-state (Hartree-Fock, HF) atom. This approximation is very inaccurate for the
positron-atom problem. It fails to describe the scattering cross sections and grossly
where 𝜑n (𝐫) ≡ 𝜑nlm (𝐫) = 1r Pnl (r)Ylm (̂𝐫 ) is the wavefunction of electron in subshell
nl. Equation (6) is equivalent to IPA. After integration over the directions of 𝐏 in the
spectrum (2), all positron partial waves contribute to the AMD incoherently. This
means that the annihilation amplitude can be calculated independently for each 𝓁,
replacing 𝜓𝐤 (𝐫) in Eq. (6) by the corresponding positron partial wave orbital 𝜓𝜀 (𝐫).
Omitting the index 𝓁, we denote such amplitude An𝜀 (𝐏).
As described below, the main corrections to the zeroth-order amplitude originate
from the electron-positron Coulomb interaction which increases the probability of
finding the electron and positron at the same point in space.
Figure 1 shows the amplitude An𝜀 (𝐏) in diagrammatic form [20, 21, 24, 49].4 The
total amplitude is depicted on left-hand side of the diagrammatic equation, with
the double line (𝜀) corresponding to incoming positron that annihilates an elec-
tron in orbital n, producing two 𝛾-rays (double-dashed line), and the circle with a
cross denoting the full annihilation vertex. The main contributions to the amplitude
are shown on the right-hand side of the equation. Diagram (a) is the zeroth-order
amplitude [IPA, Eq. (6)], diagram (b) is the first-order correction and diagram (c)
is the nonperturbative ‘virtual-positronium’ correction. This correction contains the
shaded ‘𝛤 -block’ which represents the sum of an infinite series of electron-positron
ladder diagrams shown in the lower part of Fig. 1.
The ladder diagrams represented by the 𝛤 -block are important because the
electron-positron Coulomb attraction supports bound states of the positronium (Ps)
atom. To form Ps, the energy of the incident positron needs to be greater than the
Ps-formation threshold EPs = I − 6.8 eV, where I is the ionization potential of the
atom. However, even at lower energies where the Ps can only be formed virtually, this
process gives a noticeable contribution. In practice, the 𝛤 -block is found from the
linear equation 𝛤 = V + V𝜒𝛤 , shown diagrammatically in the lower part of Fig. 1,
where V is the electron-positron Coulomb interaction and 𝜒 is the propagator of the
intermediate electron-positron state. Discretizing the electron and positron continua
by confining the system in a spherical cavity reduces this to a linear matrix equation,
which is easily solved numerically (see [19, 21, 48, 50] for further details).
4 It
is also possible to develop a diagrammatic expansion for Zef f [19, 20, 44, 45, 48] that enables
one to calculate the annihilation rate directly, rather than from Eq. (4).
248 D. G. Green and G. F. Gribakin
ν1 ν2 ν1 ν2 ν1 ν2 ν1 ν2 ν1 ν2 ν1 ν ν2
Γ ≡ + + + ... = + Γ
μ1 μ2 μ1 μ2 μ1 μ2 μ1 μ2 μ1 μ2 μ1 μ μ2
{ }
An𝜀 (𝐏) = e−i𝐏⋅𝐫 𝜓𝜀 (𝐫)𝜑n (𝐫) + 𝛥̃𝜀 (𝐫; 𝐫1 , 𝐫2 )𝜓𝜀 (𝐫1 )𝜑n (𝐫2 )d3 𝐫1 d3 𝐫2 d3 𝐫. (7)
∫
Here, the first term, corresponding to the diagram Fig. 1a, is simply the Fourier trans-
form of the product of electron and positron wavefunctions, taken at the same point.
The second term, involving the non-local annihilation kernel 𝛥̃𝜀 (of non-trivial form),
describes the vertex corrections. Note that An𝜀 (𝐏) is the Fourier transform of the cor-
related pair wavefunction (the term in the braces5 ). References [49, 50] present the
partial-wave analysis and corresponding working analytic expressions for the matrix
elements involving the vertex corrections.
As mentioned above, accurate annihilation rates and 𝛾-spectra can be obtained only
by taking into account the positron-atom correlation potential. This potential is
described by another class of diagrams that “dress” the positron wavefunction. The
corresponding positron quasiparticle wavefunction (or Dyson orbital, double line in
Fig. 1) is calculated from the Dyson equation (see, e.g., [51–53])
5 The term in braces can also be compared with the expression for the natural geminal corresponding
√
to the positron state 𝜀 and electron orbital n, 𝛼𝜀n (𝐫, 𝐫) = 𝛾𝜀n (𝐫)𝜓𝜀 (𝐫)𝜑n (𝐫) (cf. Eq. (9) in Ref. [34]),
which can be used to determine the position dependent EF 𝛾𝜀n (𝐫), see Sect. 4.
Enhancement Factors for Positron Annihilation on Valence . . . 249
Fig. 2 Main contributions to the positron self-energy matrix ⟨𝜀′ |𝛴̂ E |𝜀⟩. The lowest, second-order
diagram a describes the effect of polarization; diagram b accounts for virtual Ps formation repre-
sented by the 𝛤 -block. Diagrams c–g represent leading third-order corrections not included in (b).
Top lines in the diagrams describe the positron. Other lines with the arrows to the right are excited
electron states, and to the left, holes, i.e., electron states occupied in the target ground state. Wavy
lines represent Coulomb interactions. Summation over all intermediate states is assumed
( )
Ĥ 0 + 𝛴̂ 𝜀 𝜓𝜀 (𝐫) = 𝜀𝜓𝜀 (𝐫). (8)
Here Ĥ 0 is the Hamiltonian of the positron in the static field of the N-electron atom
in the ground state (described at the HF level). The nonlocal, energy-dependent cor-
relation potential 𝛴̂ 𝜀 is equal to the self-energy of the positron Green’s function [54],
and acts as an integral operator 𝛴̂ 𝜀 𝜓𝜀 (𝐫) = ∫ 𝛴𝜀 (𝐫, 𝐫 ′ )𝜓𝜀 (𝐫 ′ )d3 𝐫 ′ .
The main contributions to 𝛴̂ 𝜀 are shown in Fig. 2. At large positron-atom dis-
tances the correlation potential reduces to the local polarization potential 𝛴𝜀 (𝐫, 𝐫 ′ ) ≃
−𝛼d 𝛿(𝐫 − 𝐫 ′ )∕2r4 , where 𝛼d is the dipole polarizability of the atom. If only diagram
Fig. 2a is included, the polarizability is given by the HF approximation
2 ∑ |⟨𝜇|𝐫|n⟩|2
𝛼d = . (9)
3 n,𝜇 𝜀𝜇 − 𝜀n
Diagrams Fig. 2c–f are third-order corrections to the polarization diagram (a) of the
type described by the random-phase approximation with exchange [55]. Including
these gives asymptotic behaviour of 𝛴𝜀 (𝐫, 𝐫 ′ ) with a more accurate value of 𝛼d . The
diagram Fig. 2b describes the virtual Ps-formation contribution. Adding it to diagram
Fig. 2a nearly doubles the strength of the correlation potential in heavier noble-gas
atoms (Ar, Kr and Xe). The diagram Fig. 2g describes the positron-hole repulsion.
Including the diagrams of Fig. 2 in the positron-atom correlation potential provides
accurate scattering phaseshifts and cross sections for all noble-gas atoms [19].
250 D. G. Green and G. F. Gribakin
The positron self-energy diagrams and the annihilation amplitude contain sums
over the intermediate excited electron and positron states. In practice we calculate
them numerically using sets of electron and positron basis states constructed using
40 B-splines of order 6, in a spherical box of radius 30 a.u. We use an expoten-
tial knot sequence for the B-splines, which provides for an efficient spanning of the
electron and positron continua in the sums over intermediate states [48]. The max-
imum angular momentum of the intermediate states is lmax =15, and we extrapolate
to lmax → ∞ as in [19, 21, 48, 50].
Figures 3 and 4 show the AMD |An𝜀 (𝐏)|2 [spherically averaged, as in Eq. (2)] for
thermal (k = 0.04 a.u.) s-wave positrons annihilating on individual core and valence
subshells of the noble gas atoms, calculated using different approximations for the
annihilation amplitude and positron wavefunction. The range of two-𝛾 momenta P =
0–6 a.u. corresponds to the maximum Doppler energy shift 𝜖 ≈ 11 keV.
The simplest approximation shown uses the zeroth-order (IPA) annihilation ampli-
tude (6) with positron wavefunctions in the static field of the HF atom. Better approx-
imations involve using the full annihilation vertex of Fig. 1 [Eq. (7)], or the best
(Dyson) positron wavefunction, or both. In general, including correlations of either
types increases the AMD and the annihilation probability.
General trends are observed throughout the noble-gas sequence. The AMD are
broader for the core orbitals for which the typical electron momenta are greater,
leading to greater Doppler shifts. The core AMD (and the core annihilation proba-
bilities [24, 50]) also have noticeably smaller magnitudes compared with those of the
valence electrons. For all electron orbitals whose radial wavefunctions have nodes
(e.g., 2s in Ne, 2s, 3s and 3p in Ar, etc.) the AMD display deep minima related to the
nodes of the annihilation amplitude An𝜀 (𝐏). Their number and positions are related
to the number and positions of the nodes in the orbital’s radial wavefunction (i.e., the
radial nodes that occur closer to the nucleus result in the nodes of An𝜀 (𝐏) at higher
momenta). This behaviour is easy to understand from the zeroth-order amplitude (6),
which is the Fourier transform of the product of the electron and positron wavefunc-
tions. For low positron energies, its wavefunction inside the atom decreases mono-
tonically towards the nucleus (suppressed by the repulsive electrostatic potential at
smaller distances), and has no nodes. Hence, the nodal structure of the annihilation
amplitude is determined by the behaviour of the electron wavefunction. Inclusion of
the correlation corrections to the annihilation vertex, as described by Eq. (7), leads
only to a small shift in the positions of the nodes.
For a given approximation for the annihilation vertex, the AMD calculated using
the positron Dyson orbitals (red curves) are larger than those calculated using the
HF positron wavefunction (blue curves). The corresponding increase is nearly the
Enhancement Factors for Positron Annihilation on Valence . . . 251
2 2
10 10
1
1 10
10
He 1s 10
0 Ar 3p
0
10 -1
10
2
-1 -2
2
|A(P)|
10 10
|A(P)|
-3
-2 10
10
-4
-3 10
10 -5
10
-4
10 -6
10
-5 -7
10 10
1 2
10 10
1
0 10
10 Ne 2p 0 Ar 3s
10
-1 -1
10 10
2
2
-2
|A(P)|
|A(P)|
-2 10
10 -3
10
-3 -4
10 10
-5
-4 10
10 -6
10
-5 -7
10 10
1 2
10 10
1
0 10
10 Ne 2s 0 Ar 2p
10
-1 -1
10 10
2
-2
|A(P)|
|A(P)|
-2 10
10 -3
10
-3 -4
10 10
-5
-4 10
10 -6
10
-5 -7
10 10
1 2
10 10
1
0 10
10 Ne 1s 0 Ar 2s
10
-1 -1
10 10
2
-2
|A(P)|
|A(P)|
-2 10
10 -3
10
-3 -4
10 10
-5
-4 10
10 -6
10
-5 -7
10 10
0 1 2 3 4 5 6 0 1 2 3 4 5 6
P (a.u.) P (a.u.)
Fig. 3 Annihilation momentum density as a function of the total 2𝛾 momentum P, for s-wave
positron annihilation in He, Ne and Ar, calculated in different approximations for the annihilation
vertex: zeroth-order vertex (dashed lines); full vertex (0 + 1 + 𝛤 ) (solid lines), for HF (thin blue
lines) and Dyson (thick red lines) positron of momentum k = 0.04 a.u. (For a given approximation
for the vertex, the lines for the calculation with Dyson positron lie above those for HF positron)
252 D. G. Green and G. F. Gribakin
3 4
10 10
3
2 10
10 2
1 Kr 4p 10 Xe 5p
10 1
0 10
10 0
10
2
2
-1
|A(P)|
|A(P)|
10 -1
-2 10
10 -2
-3
10
10 -3
10
-4 -4
10 10
-5 -5
10 10
-6 -6
10 10
3 4
10 10
3
2 10
10 2
1 Kr 4s 10 Xe 5s
10 1
0 10
10 0
10
2
2
-1
|A(P)|
|A(P)|
10 -1
-2 10
10 -2
-3
10
10 -3
10
-4 -4
10 10
-5 -5
10 10
-6 -6
10 10
3 4
10 10
3
2 10
10 2
1 Kr 3d 10 Xe 4d
10 1
0 10
10 0
10
2
2
-1
|A(P)|
|A(P)|
10 -1
-2 10
10 -2
-3
10
10 -3
10
-4 -4
10 10
-5 -5
10 10
-6 -6
10 10
3 4
10 10
3
2 10
10 2
1 Kr 3p 10 Xe 4p
10 1
0 10
10 0
10
2
2
-1
|A(P)|
|A(P)|
10 -1
-2 10
10 -2
-3
10
10 -3
10
-4 -4
10 10
-5 -5
10 10
-6 -6
10 10
3 4
10 10
3
10
2 10
2
1 Kr 3s 10 Xe 4s
10 1
0 10
10 0
10
2
2
-1
|A(P)|
|A(P)|
10 -1
-2 10
10 10
-2
-3 -3
10 10
-4 -4
10 10
-5 -5
10 10
-6 -6
10 0 1 2 3 4 5 6
10 0 1 2 3 4 5 6
P (a.u.) P (a.u.)
Fig. 4 Annihilation momentum densities as a function of the total 2𝛾 momentum P, for s-wave
positron annihilation in Kr and Xe. Various lines as described in Fig. 3
Enhancement Factors for Positron Annihilation on Valence . . . 253
same for the valence and core orbitals in a given atom. It ranges from a factor of
∼2 in Helium to ∼100 in Xe, and is only weakly dependent on P. (Note that the
AMD are plotted on the logarithmic scale.) This increase is due to the action of the
attractive correlation potential 𝛴̂ 𝜀 on the positron (Dyson) wavefunction. It leads to
a build-up of the positron density in the vicinity of the atom and greater overlap with
the atomic electron density. This effect is stronger for the heavier, more polarizable
atoms, leading to much greater Zef f values. In Ar, Kr and Xe the attractive positron-
atom potential supports low-lying virtual s levels, leading to a characteristic resonant
growth of the annihilation rates at low positron energies [19, 44, 45, 56].
When the vertex corrections are included in the annihilation amplitude (solid
curves), the AMD is enhanced above the zeroth-order (IPA) result (dashed curves).
This is due to the Coulomb attraction within the annihilating pair, which increases
the probability of finding the electron and positron at the same point in space. The
size of the enhancement is similar for the HF and Dyson positron wavefunctions. At
the same time, the enhancement is much greater for the valence electrons than for the
core electrons, as the former are more easily perturbed by the positron’s Coulomb
field. For the core electrons, the vertex correction is dominated by the first-order
diagram Fig. 1b (similar to the case of hydrogen-like ions [21]). For the valence
electrons, the nonperturbative 𝛤 -block contribution Fig. 1c is also very important.
From Figs. 3 and 4 one can also see that the vertex enhancement is significantly
stronger at low momenta P of the electron-positron pair (which leads to narrowing
of the 𝛾-ray spectra in comparison with those obtained with the zeroth-order ampli-
tude [20, 24]). This can be seen most clearly in the AMD of the valence electrons.
Their high-P content is due to positron annihilation with the electron when the latter
is closer to the nucleus, and where its local velocity is higher, making it less suscep-
tible to the positron’s attraction. (Note that for large P the calculated 𝛤 -block vertex
corrections contain numerical errors which manifest themselves as extra oscillations
visible in AMD for valence electrons. This, however, has a negligible effect on the
annihilation spectra, since the AMD for the valence electrons at such momenta are
very small.) The vertex enhancement is considered in more detail below.
We conclude this section by noting that calculations of the corresponding 𝛾 spec-
tra were reported in [24, 50]. They showed excellent agreement with the measured
spectra for Ar, Kr and Xe [3], and firmly established the fraction of core annihilation
for these atoms.
The enhancement of the AMD and annihilation rates due to the correlation cor-
rections to the vertex with respect to those obtained using the zeroth-order (IPA)
approximation [cf. Eqs. (7) and (6)] can be parameterized through so-called vertex
enhancement factors. The MBT enables a direct ab initio calculation of ‘exact’ EF:
one simply compares the results obtained using the annihilation amplitude calculated
in different approximations, as described here.
254 D. G. Green and G. F. Gribakin
The EF were introduced originally to correct the IPA annihilation rates for
positrons in condensed matter (see, e.g., [31] and references therein),
where n− (𝐫) and n+ (𝐫) are the electron and positron densities, respectively, and
𝛾(𝐫) is the EF. The latter is typically computed for a uniform electron gas (e.g.,
using MBT [32, 57]) and parameterized in terms of the electron density, e.g., as
𝛾 = 1 + 1.23rs − 0.0742rs2 + 16 rs3 , where rs = (3∕4𝜋n− )1∕3 [58] (see also [30, 59]).
An approximation commonly used to account for the vertex enhancement of the IPA
annihilation amplitude (6) is [5, 60]
√
An𝜀 (𝐏) = e−i𝐏⋅𝐫 𝜓𝜀 (𝐫)𝜓n (𝐫) 𝛾(𝐫)d3 𝐫. (11)
∫
However, the presence of nodes in the wavefunctions in the denominator renders this
quantity of limited use and we must opt for a more pragmatic approach.
It is clear from Figs. 3 and 4 that the vertex enhancement of the AMD |An𝜀 (𝐏)|2 ,
i.e., full-vertex results compared with zeroth-order, has a weak dependence on the
momentum P (except near the nodes of the amplitude). This momentum dependence
of the vertex enhancement has little effect on the annihilation 𝛾 spectra for the noble-
gas atoms, especially for the core orbitals [50]. It is thus instructive to define a two-𝛾
momentum-averaged vertex EF as the ratio of the full-vertex partial annihilation rate
to that calculated using the zeroth-order (IPA) vertex:
Zef(0+1+𝛤
f ,nl
)
(k)
𝛾̄nl (k) = , (13)
Zef(0)f ,nl (k)
where the superscript denotes the vertex order (see Fig. 1) and nl labels the subshell
of the electron that the positron of momentum k annihilates with. Analogous EF are
commonly used to analyse and predict the annihilation rates and 𝛾 spectra in solids
Enhancement Factors for Positron Annihilation on Valence . . . 255
(see, e.g., [5, 61–66]). The true spectrum for annihilation on a given subshell for a
given positron momentum can then be approximated by
where w(0) nl
is the 𝛾 spectrum calculated using the zeroth-order vertex. Accurate
reconstruction of the true spectra for s-wave thermal positrons using Eq. (14) has
been demonstrated for noble-gas atoms in [50].
In a recent paper [24] we showed that for thermal s-wave positrons (k = 0.04 a.u.)
𝛾̄nl follows a near-universal scaling with the orbital ionization energy Inl ,
√
𝛾̄nl = 1 + A∕Inl + (B∕Inl )𝛽 , (15)
where A, B and 𝛽 are constants.6 The second term on the RHS of Eq. (15) describes
the effect of the first-order correction, Fig. 1b. Its scaling with Inl was motivated
by the 1∕Z scaling for positron annihilation in hydrogen-like ions [21]. The third
term is phenomenological and describes the effect of the 𝛤 -block correction that is
particularly important for the valence subshells.
Here we extend the calculations of the enhancement factors to s-, p- and d-wave
positrons with momenta up to the positronium-formation threshold. At small, e.g.,
room-temperature, thermal positron momenta k ∼ 0.04 a.u., the contributions of the
positron p and d waves to the annihilation rates are very small, owing to Zef f (k) ∝ k2𝓁
low-energy behaviour. (This is a manifestation of the suppression of the positron
wavefunction in the vicinity of the atom by the centrifugal potential 𝓁(𝓁 + 1)∕2r2 .)
However, for higher momenta close to the Ps-formation threshold, the s-, p- and d-
wave contributions to the annihilation rates become of comparable magnitude (see
Fig. 16 and Tables III–VII in Ref. [19]).
Figures 5–9 show the enhancement factors for positron annihilation with elec-
trons in the valence (np and ns) and core [(n − 1)s, (n − 1)p, (n − 1)d, as applicable]
orbitals of He, Ne, Ar, Kr, and Xe, as functions of the positron momentum.
The EF for positron annihilation with 1s electrons in He (Fig. 5) are 2.6–3.0 for
the s-wave, 3.8–4.1 for the p-wave, and 5.2–5.9 for the d-wave. They show only
a weak dependence on the positron momentum, which is a typical feature of all
the data. There is also little difference between the EF obtained with the static-
field (HF) positron wavefunctions (dashed lines) and those found using the positron
Dyson orbitals (solid lines). This is in spite of the fact that the use of the correlated
Dyson positron wavefunctions increases the AMD (and the annihilation rates [19])
by almost an order of magnitude for s-wave positrons (Fig. 3).
The weak dependence of the EF on the positron energy and the type of positron
wavefunction used is related to the nature of the vertex enhancement. The inter-
mediate electron and positron states in diagrams Fig. 1b and c that describe the
6
For HF positron wavefunctions the values of the parameters are A = 1.54 a.u. = 42.0 eV, B =
0.92 a.u. = 24.9 eV, and 𝛽 = 2.54. For Dyson positron wavefunctions the values are A = 1.31 a.u. =
35.7 eV, B = 0.83 a.u. = 22.7 eV, and 𝛽 = 2.15 [24].
256 D. G. Green and G. F. Gribakin
6
s-wave p-wave
5
4 d-wave
γ
0
0 0.2 0.4 0.6 0.8 1 1.2 0 0.2 0.4 0.6 0.8 1 1.2 0 0.2 0.4 0.6 0.8 1 1.2
k (a.u.) k (a.u.) k (a.u.)
Fig. 5 Enhancement factors for s-, p- and d-wave positrons annihilating on the 1s electrons in He,
obtained with HF (dashed lines) and Dyson (solid lines) positron wavefunctions
short-range vertex enhancement (𝜈, 𝜇, 𝜈1 , 𝜇1 , etc.) are highly virtual, i.e., have rel-
atively large energies. For example, the energy denominator of diagram Fig. 1b is
𝜀 − 𝜀𝜈 − 𝜀𝜇 + 𝜀n (see Refs. [20, 50]). Estimating the typical electron and positron
energies as 𝜀𝜈,𝜇 ∼ |𝜀n | (the ionization energy of electron orbital n), we see that for
few-electronvolt positrons, the positron energy 𝜀 can be neglected. For the same rea-
son, the vertex correction function 𝛥̃𝜀 (𝐫; 𝐫1 , 𝐫2 ) is only weakly nonlocal, i.e., it is large
only for |𝐫1 − 𝐫2 | ≪ |𝐫1,2 | ∼ |𝐫| (see the “annihilation maps” in Figs. 4.14–4.16 of
Ref. [67]). The situation becomes different at large momenta close to the Ps forma-
tion threshold. Here the p- and d-wave EF show an upturn related to the virtual Ps
formation becoming “more real” [68]. This is also seen in 𝛾̄np for heavier atoms.
The increase of the EF with the positron orbital angular momentum 𝓁 seen in
Fig. 5 can be related to the behaviour of the low-energy positron wavefunctions near
the atom. Due to the action of the centrifugal potential, the p- and d-wave radial
wavefunctions are suppressed as (kr)𝓁 with 𝓁 = 1 and 2, compared with the s wave.
The nonlocal correlation corrections Fig. 1b and c “help” the positron to pull the
atomic electron towards larger distances, which has a greater advantage for the higher
partial waves.
It is interesting to compare the values of 𝛾̄1s for He with the EF for positron anni-
hilation with atomic hydrogen: 6–7, 10–12, and 15–17, for the s-, p-, and d-wave
positrons, respectively, with k ≤ 0.4 a.u. (see Fig. 13 in Ref. [48]). The greater val-
ues of the EF for hydrogen are related to the smaller binding energy of the 1s electron
in hydrogen (13.6 eV) compared with that in He (24.6 eV). The vertex corrections
are generally greater for the more weakly bound electrons that have more diffuse
orbitals and are more easily perturbed by the positron’s Coulomb interaction. The
same trend will be seen throughout the noble-gas-atom sequence, with more strongly
bound electron orbitals, in particular those in the core, displaying smaller EF [cf.
Eq. (15)].
Enhancement Factors for Positron Annihilation on Valence . . . 257
5.00
s-wave p-wave
4.00 2p
d-wave
2p
3.00
2p
2s
γ
2s
2.00 2s
1s 1s 1s
1.00
0.00
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
k (a.u.) k (a.u.) k (a.u.)
Fig. 6 Enhancement factors for s-, p- and d-wave positrons annihilating on the 1s, 2s and 2p sub-
shells in Ne, obtained with HF (dashed lines) and Dyson (solid lines) positron wavefunctions.
Turning to Ne (Fig. 6), we observe that the EF for the outer valence 2p subshell
are slightly smaller than those for 1s in He, in spite of the binding energy of the 2p
electrons (21.6 eV) being lower than that of He 1s. Ne also has the broadest 𝛾 ray
spectrum of all the noble gases (see AMD in Fig. 3, and the data for the calculated
and measured spectra [50, 69]). The latter indicates that the 2p electrons in Ne have
large typical momenta, which makes the correlation correction to the annihilation
vertex relatively small. The EF for the inner valence 2s subshell is around 2, while
for the deeply bound 1s electrons, 𝛾̄1s ≈ 1.2. We also note that for the core orbitals,
the values of the EF for the positron s, p and d waves are quite close. This is in fact
a general trend observed for all atoms that the relative difference between the values
of 𝛾̄nl − 1 for the positron s, p and d waves is becoming small with the increase in the
binding energy. The smaller effect of the orbital angular momentum of the positron
on the EF for core orbitals is due to the vertex correction becoming “more local”,
and hence, less sensitive to the variation of the positron radial wavefunction.
The EF in Ar, Kr and Xe (Figs. 7, 8 and 9) become progressively larger, for
both the valence and core electrons. For example, the vertex EF for s-wave positron
annihilation with the outer valence np electrons increases from 𝛾̄3p = 5.2 (Ar), to
𝛾̄4p = 6.6 (Kr), to 𝛾̄5p = 9.2 (Xe) (for the HF positron wavefunction at low momenta
k ≲ 0.1 a.u.). The EF for the (n − 1)l core orbitals also increase to 𝛾̄(n−1)l ∼ 1.5–2,
with the values for the 3d and 4d orbitals being noticeably larger than those of the
3s/3p and 4s/4p orbitals, for Kr and Xe, respectively.
Another feature of the data is the growing difference between the EF for the np
electrons obtained with the Dyson positron wavefunction (solid lines) and those
found using the static-field (HF) positron wavefunction (dashed lines). This effect
is related to the increase in the strength of the positron-atom correlation potential
𝛴̂ 𝜀 for the heavier noble-gas atoms [19, 44, 45]. For s-wave positrons it results in
the creation of positron-atom virtual levels [47] whose energies 𝜀 = 𝜅 2 ∕2 become
lower for heavier atoms, with values of 𝜅 = −0.23, −0.10 and −0.012 a.u. for Ar,
258 D. G. Green and G. F. Gribakin
10.00
9.00 s-wave p-wave d-wave
8.00 3p
7.00
3p
6.00
5.00 3p
γ
4.00
3s 3s
3.00 3s
2.00 2s, 2p 2s, 2p 2s,2p
1.00
0.00
0 0.2 0.4 0.6 0.8 0 0.2 0.4 0.6 0.8 0 0.2 0.4 0.6 0.8
k (a.u.) k (a.u.) k (a.u.)
Fig. 7 Enhancement factors for s-, p- and d-wave positrons annihilating on the 2s, 2p, 3s and 3p
subshells in Ar, obtained with HF (dashed lines) and Dyson (solid lines) positron wavefunctions
12.00
8.00
4p
6.00
γ
4p
4.00 4s
4s 4s
2.00 3d 3d 3d
0.00
0 0.2 0.4 0.6 0 0.2 0.4 0.6 0 0.2 0.4 0.6
k (a.u.) k (a.u.) k (a.u.)
Fig. 8 Enhancement factors for s-, p- and d-wave positrons annihilating on the 3s, 3p, 3d, 4s and
4p subshells in Kr, obtained with HF (dashed lines) and Dyson (solid lines) positron wavefunctions
Kr and Xe, respectively [19]. This is accompanied by a rapid growth of the positron
wavefunction near the atom, with the Dyson orbitals being enhanced by a factor
∼1∕|𝜅| compared to the static-field positron wavefunctions at low energies. Hence,
the inclusion of the correlation potential makes the radial dependence of the positron
wavefunction more vigorous. This is evidenced by some broadening of the 𝛾 spectra
obtained with the Dyson rather than the HF positron wavefunction [50]. This also
Enhancement Factors for Positron Annihilation on Valence . . . 259
18
16
s-wave p-wave d-wave
14
12 5p
10
γ
5p
8
5p
6
5s
4 5s 5s
4d 4d 4d
2
4s, 4p 4s, 4p 4s, 4p
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5
k (a.u.) k (a.u.) k (a.u.)
Fig. 9 Enhancement factors for s-, p- and d-wave positrons annihilating on the 4s, 4p, 4d, 5s and
5p subshells in Xe, obtained with HF (dashed lines) and Dyson (solid lines) positron wavefunctions
results in a reduction of the EF, which is most noticeable for the valence electrons,
and is largest in Xe, which has the strongest correlation potential for the positron.
Besides the rise in the valence EF for high positron momenta (related to the prox-
imity of the Ps-formation threshold), one other exception from the weak momentum-
dependence of the EF is seen at low momenta for d-wave positron annihilating on the
3d and 4d orbitals in Kr and Xe. The enhancement factors in this case are approxi-
mately constant from the Ps-formation threshold down to ∼0.3 a.u., but then deviate
at lower momenta, especially in Xe. Both the zeroth-order and full-vertex Zef f values
for these orbitals are calculated to be smooth functions of k. However, they obey the
∼k4 behaviour and become very small at low k (e.g., for Xe, using the Dyson wave-
function we find Zef f,4d ∼ 10−3 at k ∼ 0.3 a.u., decreasing to ∼10−7 for k ∼ 0.03 a.u.
It appears that for such small k numerical inaccuracies arise in the calculation of the
𝛤 -block contribution, leading to errors when evaluating the ratio in Eq. (13).
5 Conclusions
increasing electron binding energy. We find that the type of the positron wavefunc-
tion used, i.e., Dyson orbital which accounts for the positron-atom correlation attrac-
tion vs repulsive static-field (HF) wavefunction, has relatively little effect on the EF,
except for the valence orbitals in most polarisable targets. We also find a relatively
weak dependence of the EF for the core and inner-valence electrons on the positron’s
orbital angular momentum.
The weak momentum-dependence of the EF obtained in positron-atom calcula-
tions suggests that they can be used to improve the calculations of positron anni-
hilation in more complex environments. One such system is positronium colliding
with noble-gas atoms, where calculations of Ps-atom pick-off annihilation rates that
neglect the short-range vertex enhancement strongly underestimate the measured
rates [70]. Another context where similar EF can be used is positron annihilation in
molecules. Here there is a sharp contrast between the large amount of experimen-
tal information, including 𝛾-spectra, for a wide range of molecule [69] and paucity
of credible theoretical data [23, 71]. The positron-molecule problem is particularly
interesting because the Zeff values for most polyatomic molecules show orders-of-
magnitude increases due to resonant positron annihilation [71]. In such molecules
the positron annihilates from a temporarily formed weakly-bound state. Attempts to
calculate such states using standard quantum-chemistry methods have been numer-
ous but not very successful [71] (i.e., there is only a small number of systems where
theory and experiment can be compared, and the agreement is mostly qualitative
[72]).
The calculations presented in this paper could be extended to other atoms, includ-
ing those with open shells. Partial filling of electron shells can be taken into account
in the many-body-theory sums using fractional occupation numbers (cf. Ref. [73]),
and the positron wavefunction is insensitive to such details of the electronic
structure at the static (HF) level. (At the level of Dyson orbitals, one will need
to calculate the positron self-energy using fractional electron-shell occupancies.)
Calculations are particularly straightforward for annihilation with core electrons, as
their enhancement factor is described well by the first-order correction to the anni-
hilation vertex.
References
1. Asoka-Kumar P, Alatalo M, Ghosh V, Kruseman A, Nielsen B, Lynn K (1996) Phys Rev Lett
77:2097. https://doi.org/10.1103/PhysRevLett.77.2097
2. Lynn KG, MacDonald JR, Boie RA, Feldman LC, Gabbe JD, Robbins MF, Bonderup E,
Golovchenko J (1977) Phys Rev Lett 38:241. https://doi.org/10.1103/PhysRevLett.38.241
3. Iwata K, Gribakin GF, Greaves RG, Surko CM (1997) Phys Rev Lett 79:39. https://doi.org/10.
1103/PhysRevLett.79.39
Enhancement Factors for Positron Annihilation on Valence . . . 261
4. Lynn K, Dickman J, Brown W, Robbins M, Bonderup E (1979) Phys Rev B 20:3566. https://
doi.org/10.1103/PhysRevB.20.3566
5. Alatalo M, Barbiellini B, Hakala M, Kauppinen H, Korhonen T, Puska M, Saarinen K, Hauto-
järvi P, Nieminen R (1996) Phys Rev B 54:2397. https://doi.org/10.1103/PhysRevB.54.2397
6. Tuomisto F, Makkonen I (2013) Rev Mod Phys 85:1583. https://doi.org/10.1103/RevModPhys.
85.1583
7. Weiss A, Mayer R, Jibaly M, Lei C, Mehl D, Lynn KG (1988) Phys Rev Lett 61:2245. https://
doi.org/10.1103/PhysRevLett.61.2245
8. Ohdaira T, Suzuki R, Mikado T, Ohgaki H, Chiwaki M, Yamazaki T (1997) Appl Surf Sci
116:177. https://doi.org/10.1016/S0169-4332(96)01049-5
9. Weiss AH, Fazleev NG, Nadesalingam MP, Mukherjee S, Xie S, Zhu J, Davis BR (2007) Radiat
Phys Chem 76:285. https://doi.org/10.1016/j.radphyschem.2006.03.053
10. Mayer J, Hugenschmidt C, Schreckenbach K (2010) Surf Sci 604:1772. https://doi.org/10.
1016/j.susc.2010.07.003
11. Hugenschmidt C (2016) Surf Sci Rep 71:547. https://doi.org/10.1016/j.surfrep.2016.09.002
12. Hugenschmidt C, Lwe B, Mayer J, Piochacz C, Pikart P, Repper R, Stadlbauer M, Schreck-
enbach K (2008) Nucl Instrum Methods A 593:616. https://doi.org/10.1016/j.nima.2008.05.
038
13. Mayer J, Hugenschmidt C, Schreckenbach K (2010) Phys Rev Lett 105:207401. https://doi.
org/10.1103/PhysRevLett.105.207401
14. Stoll H, Koch KMM, Major J (1991) Nucl Instrum Methods B 582:56
15. Coleman P (ed) (2000) Positron beams and their applications. World Scientific
16. Sano Y, Kino Y, Oka T, Sekine T (2015) J Phys Conf Ser 618:012010. http://stacks.iop.org/
1742-6596/618/i=1/a=012010
17. Jensen KO, Weiss A (1990) Phys Rev B 41:3928. https://doi.org/10.1103/PhysRevB.41.3928
18. Iwata K, Greaves RG, Murphy TJ, Tinkle MD, Surko CM (1995) Phys Rev A 51:473. https://
doi.org/10.1103/PhysRevA.51.473
19. Green DG, Ludlow JA, Gribakin GF (2014) Phys Rev A 90:032712. https://doi.org/10.1103/
PhysRevA.90.032712
20. Dunlop LJM, Gribakin GF (2006) J Phys B 39:1647. https://doi.org/10.1088/0953-4075/39/7/
008
21. Green DG, Gribakin GF (2013) Phys Rev A 88:032708. https://doi.org/10.1103/PhysRevA.88.
032708
22. Green DG, Saha S, Wang F, Gribakin GF, Surko CM (2010) Mater Sci Forum 666:21. https://
doi.org/10.4028/www.scientific.net/MSF.666.21
23. Green DG, Saha S, Wang F, Gribakin GF, Surko CM (2012) New J Phys 14:035021. http://
stacks.iop.org/1367-2630/14/i=3/a=035021
24. Green DG, Gribakin GF (2015) Phys Rev Lett 114:093201. https://doi.org/10.1103/
PhysRevLett.114.093201
25. Bonderup E, Andersen JU, Lowy DN (1979) Phys Rev B 20:883. https://doi.org/10.1103/
PhysRevB.20.883
26. Flambaum VV (2009) Phys Rev A 79:042505. https://doi.org/10.1103/PhysRevA.79.042505
27. Kozlov MG, Flambaum VV (2013) Phys Rev A 87:042511. https://doi.org/10.1103/PhysRevA.
87.042511
28. Kahana S (1963) Phys Rev 129:1622. https://doi.org/10.1103/PhysRev.129.1622
29. Carbotte JP (1967) Phys Rev 155:197. https://doi.org/10.1103/PhysRev.155.197
30. Boroński E, Nieminen R (1986) Phys Rev B 34:3820. https://doi.org/10.1103/PhysRevB.34.
3820
31. Puska MJ, Nieminen RM (1994) Rev Mod Phys 66:841. https://doi.org/10.1103/RevModPhys.
66.841
32. Arponen J, Pajanne E (1979) Ann Phys 121:343. https://doi.org/10.1016/0003-
4916(79)90101-5
33. Mitroy J, Barbiellini B (2002) Phys Rev B 65:235103. https://doi.org/10.1103/PhysRevB.65.
235103
262 D. G. Green and G. F. Gribakin
34. Makkonen I, Ervasti MM, Siro T, Harju A (2014) Phys Rev B 89:041105. https://doi.org/10.
1103/PhysRevB.89.041105
35. Daniuk S, Kontrym-Sznajd G, Rubaszek A, Stachowiak H, Mayers J, Walters PA, West RN
(1987) J Phys F: Metal Phys 17:1365. https://doi.org/10.1088/0305-4608/17/6/011
36. Jarlborg T, Singh AK (1987) Phys Rev B 36:4660. https://doi.org/10.1103/PhysRevB.36.4660
37. Rubaszek A , Stachowiak H (1984) Physica Status Solidi (B) 124:159. https://doi.org/10.1002/
pssb.2221240117
38. Zubiaga A, Ervasti MM, Makkonen I, Harju A, Tuomisto F, Puska MJ (2016) J Phys B: At
Mol Opt Phys 49:064005
39. Swann AR, Green DG, Gribakin GF arXiv: 1709.00394
40. Berestetskii VB, Lifshitz EM, Pitaevskii LP (1982) Quantum electrodynamics, 2nd edn. Perg-
amon, Oxford
41. Kaijser P, Smith Jr VH (1977) Adv Quantum Chem 10:37. https://doi.org/10.1016/S0065-
3276(08)60578-X
42. Fraser PA (1968) Adv At Mol Phys 4:63
43. Pomeranchuk I, Eksp Zh (1949) Teor Fiz 19:183
44. Dzuba VA, Flambaum VV, King WA, Miller BN, Sushkov OP (1993) Phys Scripta T46:248.
https://doi.org/10.1088/0031-8949/1993/T46/039
45. Dzuba VA, Flambaum VV, Gribakin GF, King WA (1996) J Phys B 29:3151. https://doi.org/
10.1088/0953-4075/29/14/024
46. Surko CM, Gribakin GF, Buckman SJ (2005) J Phys B 38:R57. https://doi.org/10.1088/0953-
4075/38/6/R01
47. Landau LD, Lifshitz EM (1977). Quantum mechanics (Non-relativistic theory). In: Course of
theoretical physics, 3rd edn, vol 3. Pergamon, Oxford
48. Gribakin GF, Ludlow J (2004) Phys Rev A 70:032720. https://doi.org/10.1103/PhysRevA.70.
032720
49. Green DG (2011) PhD thesis, Queen’s University Belfast
50. Green DG, Gribakin GF (2015) arXiv:1502.08045
51. Abrikosov AA, Gorkov LP, Dzyalonshinkski IE (1975) Methods of quantum field theory in
statistical physics. Dover, New York
52. Fetter AL, Walecka JD (2003) Quantum theory of many-particle systems. Dover, New York
53. Dickhoff WH, Neck DV (2008) Many body theory exposed!—Propagator description of quan-
tum mechanics in many-body systems, 2nd edn. World Scientific, Singapore
54. Bell JS, Squires EJ (1959) Phys Rev Lett 3:96. https://doi.org/10.1103/PhysRevLett.3.96
55. Amusia MY, Cherepkov NA (1975) Case studies in atomic physics 5:47
56. Goldanski VI, Sayasov YS (1968) Phys Lett 13:300
57. Arponen J, Pajanne E (1979) J Phys F 9:2359. https://doi.org/10.1088/0305-4608/9/12/009
58. Barbiellini B, Puska MJ, Torsti T, Nieminen RM (1995) Phys Rev B 51:7341. https://doi.org/
10.1103/PhysRevB.51.7341
59. Stachowiak H, Lach J (1993) Phys Rev B 48:9828. https://doi.org/10.1103/PhysRevB.48.9828
60. Daniuk S, Kontrym-Sznajd G, Rubaszek A, Stachowiak H, Mayers J, Walters PA, West RN
(1987) J Phys F 17:1365. https://doi.org/10.1088/0305-4608/17/6/011
61. Jensen KO (1989) J Phys Condens Matter 1:10595. https://doi.org/10.1088/0953-8984/1/51/
027
62. Alatalo M, Kauppinen H, Saarinen K, Puska MJ, Mäkinen J, Hautojärvi P, Nieminen RM
(1995) Phys Rev B 51:4176. https://doi.org/10.1103/PhysRevB.51.4176
63. Barbiellini B, Puska MJ, Alatalo M, Hakala M, Harju A, Korhonen T, Siljamäki S, Torsti T,
Nieminen RM (1997) Appl Surf Sci 116:283. https://doi.org/10.1016/S0169-4332(96)01070-
7
64. Barbiellini B, Hakala M, Puska MJ, Nieminen RM, Manuel AA (1997) Phys Rev B 56:7136.
https://doi.org/10.1103/PhysRevB.56.7136
65. Makkonen I, Hakala M, Puska MJ (2006) Phys Rev B 73:035103. https://doi.org/10.1103/
PhysRevB.73.035103
Enhancement Factors for Positron Annihilation on Valence . . . 263
66. Kuriplach J, Morales AL, Dauwe C, Segers D, Šob M (1998) Phys Rev B 58:10475. https://
doi.org/10.1103/PhysRevB.58.10475
67. Ludlow J (2003) PhD thesis, Queen’s University Belfast
68. Gribakin GF, Ludlow J (2002) Phys Rev Lett 88:163202. https://doi.org/10.1103/PhysRevLett.
88.163202
69. Iwata K, Greaves RG, Surko CM (1997) Phys Rev A 55:3586. https://doi.org/10.1103/
PhysRevA.55.3586
70. Mitroy J, Ivanov IA (2001) Phys Rev A 65:012509. https://doi.org/10.1103/PhysRevA.65.
012509
71. Gribakin GF, Young JA, Surko CM (2010) Rev Mod Phys 82:2557. https://doi.org/10.1103/
RevModPhys.82.2557
72. Tachikawa M (2014) J Phys Conf Ser 488:012053. https://doi.org/10.1088/1742-6596/488/1/
012053
73. Dzuba VA, Kozlov A, Flambaum VV (2014) Phys Rev A 89:042507. https://doi.org/10.1103/
PhysRevA.89.042507
Geometric Phase and Interference Effects
in Ultracold Chemical Reactions
N. Balakrishnan (✉)
Department of Chemistry, University of Nevada, Las Vegas, NV 89154, USA
e-mail: [email protected]
B. K. Kendrick
Los Alamos National Laboratory, Theoretical Division (T-1, MS B221),
Los Alamos, NM 87545, USA
1 Introduction
sections [14–23]. Thus, it appears that an energy regime where only a single par-
tial wave contributes is the most relevant regime to explore GP effects in a chemical
reaction. This regime, referred to as the cold and ultracold regime, has gained much
interest in recent years, thanks to the dramatic progress in cooling and trapping of
molecules in the mK and 𝜇K regimes. Here, we will focus on our recent studies
of the GP effect in chemical reactions in the ultracold regime taking the hydrogen
exchange reaction as an illustrative example.
The ultracold regime [34–38] provides a fascinating domain to explore quan-
tum effects in chemical reactions. Because s-wave scattering dominates at ultracold
temperatures (for bosons and distinguishable particles), only the l = 0 partial wave
contributes and the GP effect is not smeared out by partial wave summation. Further-
more, isotropic scattering in the s-wave regime allows for maximum constructive or
destructive interference between wave functions along alternative paths around the
CI (direct and looping/exchange paths). These properties combined with an effec-
tive quantization of the scattering phase-shift in the ultracold regime (Levinson’s
theorem [39] 𝛿(0) = n𝜋 where 𝛿 is the phase shift and n is the number of bound
states supported by the potential well) entail maximum constructive or destruc-
tive interference between the direct and exchange/looping scattering amplitudes.
This leads to a large enhancement or suppression of reactivity, as recently demon-
strated for O+OH(v = 0, 1) → H+O2 (v′ , j′ ) [40, 41] and the hydrogen exchange
processes in H+H2 (v = 4, j = 0), H+HD(v = 4, j = 0) and D+HD(v = 4, j = 0) reac-
tions [42–45]. The H+H2 reaction has an energy barrier for vibrational levels v < 3
but becomes barrierless for v > 3 [43, 46–48]. Indeed, vibrationally adiabatic poten-
tials for the H+H2 reaction for v = 4 and higher vibrational levels depict an effective
potential well. The bound state structure of this potential well has a dramatic effect
on the scattering process at ultracold temperatures as discussed below. Also, barrier-
less reactions occur with much larger rate coefficients at ultracold temperatures and
are more amenable to experiments than barrier reactions that proceed via tunneling.
The chapter is organized as follows. In Sect. 2 we briefly discuss the mecha-
nism of the GP effect in ultracold reactions. Section 3 outlines the coupled channel
method employed in the scattering calculations. Illustrative results of GP effects in
H+H2 /H+HD/D+HD reactions are presented in Sect. 4 followed by conclusions in
Sect. 5.
In our previous work [40, 42] we showed that to observe the GP effect in reactive
and inelastic collisions two criteria should be satisfied: (i) the adiabatic PES must
exhibit a conical intersection; (ii) the scattering amplitudes along the two scattering
pathways (direct and exchange/looping) must have comparable magnitude and scat-
ter into the same angular region. Isotropic scattering in the ultracold regime and the
effective quantization of the scattering phase shift as required by Levinson’s theorem
provide the criterion for maximum constructive and destructive interference between
268 N. Balakrishnan and B. K. Kendrick
10-11 10-11
(a) (b)
v = 0, j = 8
`
v = 1, j = 4
`
10-12 10-12
10-13
Rate Coefficient (cm3/s)
10-13
10-11 v = 3, j = 4
`
10-12
10-12
10-13
10-13 J = 0-4
NGP J = 0-4 NGP
GP GP
10-14 -6 10-14 -6
10 10-4 10-2 100 102 10 10-4 10-2 100 102
Energy (K) Energy (K)
the two scattering amplitudes [40–42]. The details of the interference mechanism
have been discussed in prior works [40–42] and only a brief description is given
here.
In H+HD collisions, due to the presence of two identical H atoms, a correct treat-
ment of the H+HD channel should include both purely non-reactive collisions such
as Ha +Hb D(v, j)→Ha +Hb (v′ , j′ )D (where a and b labels the two hydrogen atoms for
illustrative purpose) and exchange collisions such as Ha +Hb D(v, j)→Hb +Ha D(v′ , j′ )
where the two identical H atoms exchange with one another. The same applies to the
D+HD system where the identical D atoms are exchanged in the exchange scatter-
ing amplitude. Both reactions also include purely reactive channels leading to D+H2
and H+D2 products. These reactive pathways may involve a “direct” path (traverses
over one transition state) or a “looping path” (traverses over two transition states).
For a schematic illustration of these reaction pathways, see Fig. 1a, b of Kendrick et
al. [42]. Our recent studies have shown large GP effects in the H+HD and D+HD
channels due to strong interference between the inelastic and exchange components
of the scattering amplitudes. The GP effect was found to be not significant for the
purely reactive channels due to small values of the scattering amplitudes for the loop-
ing pathway resulting in negligible interference with the dominant “direct” pathway.
Geometric Phase and Interference Effects in Ultracold Chemical Reactions 269
Here, we focus on the exchange channel that shows the largest GP effect. Our the-
oretical description is based on the exchange pathways depicted in Fig. 1a of Ref.
[42] and earlier works of Althorpe and collaborators [20–23].
For H+HD and D+HD systems (or in general, A+AB collisions), the GP and
NGP (no geometric phase) scattering amplitudes can be written in terms of the
“inelastic” and “exchange” scattering amplitudes, finel and fex , respectively,
1
fNGP∕GP = √ (finel ± fex ). (1)
2
The square modulus of the scattering amplitudes for the NGP and GP calculations
may be written as
1
|fNGP∕GP |2 = (|f |2 + |fex |2 ± 2|finel | |fex | cos Δ), (2)
2 inel
where the complex scattering amplitudes finel = |finel |ei𝛿inel and fex = |fex |ei𝛿ex and
Δ = 𝛿ex − 𝛿inel is the phase difference between the exchange and inelastic pathways.
For comparable values of the two scattering amplitudes, i.e., |fex | = |finel | = |f |,
Eq. (2) becomes |fNGP∕GP |2 = |f |2 (1 ± cos Δ). Furthermore, if cos Δ = +1 then max-
imum (constructive) interference occurs for the NGP case and |fNGP |2 ∼ 2|f |2 and
|fGP |2 ∼ 0. On the other hand, if cos Δ = −1 then maximum (constructive) inter-
ference occurs for the GP case and |fGP |2 ∼ 2|f |2 and |fNGP |2 ∼ 0. Recalling that
Δ = n𝜋 can occur in the ultracold regime (Levinson’s theorem) where n is an inte-
ger, the reaction can be turned on or off depending simply on the sign of the inter-
ference term (since | cos Δ| ∼ 1). In contrast, if one of the scattering amplitudes is
much greater than the other, |fex |2 ≫ |finel |2 or |finel |2 ≫ |fex |2 , then Eq. (2) becomes
|fNGP∕GP |2 ∼ |fex |2 ∕2 or |fNGP∕GP |2 ∼ |finel |2 ∕2. The GP effect vanishes in this case
and the interference term containing | cos Δ| plays no role. In the high partial wave
limit (high collision energies), the interference term averages out to zero (cos Δ ∼ 0)
and there is no GP effect. This description is also valid for the pure reactive case,
e.g., H+HD→D+H2 except the two scattering amplitudes for the different paths are
replaced by |fex | = |floop | and |finel | = |fdirect |. In our previous work, we have shown
that the phase quantization of Δ = n𝜋 can be understood by considering scattering in
a simple spherical well potential for the different pathways (i.e., Levinson’s theorem
𝛿ex = nex 𝜋 and 𝛿inel = ninel 𝜋 but with a different number of bound states nex and ninel
for the spherical well potentials traversed by the two pathways) [40].
The reactive scattering calculations were carried out using hyperspherical coor-
dinates. Two sets of hyperspherical coordinates are employed: the adiabatically-
adjusting principle axis hyperspherical (APH) coordinates of Pack and Parker
270 N. Balakrishnan and B. K. Kendrick
[49, 50] in the inner hyper-radial region where the three-body interaction is strong
and the Delves hyperspherical coordinates in the outer region where the three-body
forces vanish and different atom-diatom configurations emerge. The APH coordi-
nates are independent of the different atom-diatom arrangement channels and allow
an evenhanded description of all three arrangement channels in an A+BC system
compared to the Delves hyperspherical coordinates. The method accurately treats the
body-frame Eckart singularities [50] associated with non-zero total angular momen-
tum quantum number J and includes the geometric phase using the general vector
potential approach [15, 16, 18]. The geometric phase is included only in the APH
coordinates as it is relevant only in the region of three-body interaction where the
CI is located. Regardless of the choice of the hyperspherical coordinates, the basic
numerical approach involves a sector-adiabatic formalism. The hyper radius (𝜌) is
divided into a large number of sectors and at the center of each sector, the total wave
function is expanded in terms of five-dimensional hyperspherical surface functions.
The surface functions are in turn expanded in primitive angular functions. Conver-
gence is sought with respect to the number of primitive functions included in the
expansion. A sequential truncation/diagonalization procedure is used to reduce the
size the surface function matrix. The expansion coefficients depend on the hyper
radius but within a sector they are assumed to be independent of 𝜌. Coupled channel
equations resulting from the Schrödinger equation with this expansion of the total
wave function in terms of hyperspherical surface functions are solved from sector-
to-sector. Asymptotic boundary conditions are applied in Jacobi coordinates at the
last sector in 𝜌 to evaluate the reactance and scattering matrices from which cross
sections and rate coefficients are computed using standard expressions [49].
4 Results
d𝜎 | k̄ v′ j′ N
| = |f ′ ′ ′ − (−1) fvjm→v′ j′ m′ |2 ,
igp R
(3)
dΩ |vjm→v′ j′ m′ k̄ vj vjm→v j m
Geometric Phase and Interference Effects in Ultracold Chemical Reactions 271
where f N and f R are the scattering amplitudes for the non-reactive and reactive chan-
nels, k̄ vj are the appropriately normalized wave vector magnitudes, and igp = 1 or 0
for calculations which include or do not include the GP, respectively (see Ref. [18] for
details). As demonstrated in previous [18] studies using the vector potential approach
[1] the GP effect is captured almost entirely by the sign change (i.e., igp = 1) given
in Eq. (3). Thus, the GP can be accurately included for H + H2 by performing cal-
culations without the vector potential [18, 51]. The computed f N and f R are then
properly combined using Eq. (3) to include (igp = 1) or not include (igp = 0) the GP.
We used this approach for the H + H2 calculations reported in this work. For H+HD
and D+HD reactions, the GP effect was included using the vector potential approach.
The H3 PES of Boothroyd et al. [52] is used in the calculations reported here. We
have verified that the H3 PES of Mielke et al. [53] yields comparable results [42].
Figure 1 shows the rotationally resolved reaction rate coefficients for the
H+H2 (v = 4, j = 0) → H+H2 (v′ , j′ ) reaction for v′ = 0, j′ = 8, v′ = 1, j′ = 4, v′ =
2, j′ = 4 and v′ = 3, j′ = 4 [45]. It is seen that the GP rates dominate the NGP rates
for all four rotational levels shown in Fig. 1. Indeed, a similar trend is found for all
the state-to-state rotational transitions leading to even rotational levels of H2 for all v′
levels (see Table I of Kendrick et al. [45]). As a result, a similar GP effect is found for
vibrationally resolved rate coefficients for para-para transitions when summed over
all rotational levels in a given vibrational state. The trend also prevails in the total
rate coefficients as illustrated in Fig. 2. The upper panel of Fig. 2 shows the total rate
coefficient and the lower panel displays contributions from different partial waves
(orbital angular momentum l = J for initial rotational level j = 0).
Figure 3 shows a comparison between GP and NGP rate coefficients for the
H+HD(v = 4, j = 0) → HD(v′ , j′ )+ H reaction for v′ = 0, j′ = 3, v′ = 1, j′ = 2, and
272 N. Balakrishnan and B. K. Kendrick
`
10-15 v = 0, j = 3
`
10-15
10-16
Rate Coefficient (cm3/s)
10-16
`
`
`
10-16 10-16
10-12
Evn Ex-Sym (c) 10-12 Odd Ex-Sym (f)
10-13
10-13
10-14 10-14 v = 3, j = 2
`
v = 3, j = 2
`
10-15 10-15
10-16 10-1610-6
10-6 10-4 10-2 100 102 10-4 10-2 100 102
Energy (K) Energy (K)
Fig. 3 Rotationally resolved rate coefficients (cross section times the relative velocity) for the
H+HD(v = 4, j = 0) → H+HD(v′ , j′ ) reaction for v′ = 0, j′ = 3, v′ = 1, j′ = 2, and v′ = 3, j′ = 2 as
functions of the collision energy. Even and odd exchange symmetry results are given in the left
and right panels, respectively. In each panel, the red curves show the GP results and the black
curves denote the NGP results. The results include all values of total angular momentum J = 0–4.
Reproduced with permission from [43]
10-10
Evn+Odd Symm
NGP
GP
Rate Coefficient (cm3/s) 10-12
10-14 J = 0-4
J=0
J=1
J=2
J=3
10-16 J=4
10-18 -7
10 10-6 10-5 10-4 10-3 10-2 10-1 100 101 102
Energy (K)
Fig. 4 Total angular momentum resolved (partial wave resolved since l = J for j = 0) rate coef-
ficients for the H+HD(v = 4, j = 0) → H+HD reaction summed over all energetically accessible
rotational and vibrational levels of the product HD molecule as well as even and odd exchange
symmetry contributions weighted by appropriate nuclear spin statistics factors. Reproduced with
permission from [43]
of the scattering amplitudes for the exchange and direct paths (upper panel) and the
⟨cos Δ⟩ values (lower panel) as a function of the collision energy for the v′ = 0, j′ =
3 transition shown in Fig. 3 for the H+HD reaction. It is seen that the NGP rates
dominate when ⟨cos Δ⟩ = +1 and the GP rates dominate when ⟨cos Δ⟩ = −1. This
is most clearly seen at energies below 1 mK where s-wave scattering dominates.
At higher collision energies where non-zero partial waves contribute, the ⟨cos Δ⟩
values oscillate around zero making the contribution from the interference term in
Eq. (2) less significant leading to a negligible GP effect. This trend is observed in all
state-to-state rotationally resolved rate coefficients for H+HD and D+HD reactions
[43, 44].
The large GP effect in state-to-state cross sections when isotropic scattering dom-
inates in the s-wave regime is illustrated in Fig. 6 where differential cross sections
(DCSs) for the v′ = 3, j′ = 0 final state in the D+HD(v = 4, j = 0) reaction are plot-
ted as a function of the collision energy and the scattering angle. It is seen that the
NGP results are enhanced for even exchange symmetry while the GP results are
enhanced for the odd exchange symmetry through constructive interference between
the scattering amplitudes for the direct and exchange pathways. The interference is
destructive when the DCSs for these cases are suppressed. The interference pattern
changes when resonances are present as seen at energies near 1 K where a l = 2
shape resonance occurs for the D+HD reaction. A detailed discussion of the reso-
nances and parameters characterizing them (position, width and lifetimes) are given
in Hazra et al. [43] and Kendrick et al. [44] for the H+HD and D+HD reactions.
274 N. Balakrishnan and B. K. Kendrick
1.4
Evn Ex-Sym (a) Odd Ex-Sym (c)
1.3 1.35
v = 0, j = 3
`
1.2 1.3 J=0
J=1
v = 0, j = 3
`
J=0 J = 0-4
1.1 J=1
J = 0-4 1.25
1.5 1.5
Evn Ex-Sym (b) Odd Ex-Sym (d)
1 1
v = 0, j = 3
`
0.5 0.5
v = 0, j = 3
`
`
COS Δ
COS Δ
0 0
J=0 J=0
J=1 J=1
-0.5 J = 0-4 -0.5 J = 0-4
-1 -1
-1.5 -1.5
10-6 10-4 10-2 100 102 10-6 10-4 10-2 100 102
Energy (K) Energy (K)
Fig. 5 Ratio of the square of the scattering amplitudes for the exchange (reactive) and inelastic
(non-reactive) pathways (upper panels) and average value of cos Δ (lower panels) for both even (left
panels) and odd (right panels) exchange symmetries as functions of the collision energy. Results are
presented for the H+HD(v′ = 0, j′ = 3) product channel. Reproduced with permission from [43]
We have discussed the importance of the geometric phase effect in ultracold hydro-
gen exchange reactions in collisions of H and D atoms with vibrationally excited
H2 and HD molecules. For vibrational levels v > 3 these reactions occur through
a barrierless path. Results presented for the v = 4 vibrational levels of the H2 and
HD molecules illustrate strong interference between the direct and exchange com-
ponents of the scattering amplitudes leading to enhancement or suppression of the
reactivity. Isotropic scattering in the ultracold s-wave regime allows maximum con-
structive/destructive interference leading to large GP effects in state-to-state reaction
rate coefficients. The effect persists but to a lesser extent in the total reaction rate, due
in part, to the cancellation of the GP effects when even and odd exchange symmetry
results are added to yield the total rates for H+HD and D+HD reactions. The results
presented here illustrate that the GP effect may be experimentally observable by the
selection of a particular nuclear spin-state of the HD molecule in H+HD/D+HD
collisions.
Geometric Phase and Interference Effects in Ultracold Chemical Reactions 275
180
(deg)
140
GP
100
angle
60
ering
scatt
20
log(Ec) K
180
(deg)
140
NGP
100
ngle
ring a
60
e
scatt
20
log(Ec) K
Fig. 6 The DCS is plotted as a function of collision energy and scattering angle for the D +
HD(v = 4, j = 0) → D + HD(v′ = 3, j′ = 0) reaction. Panel a is even exchange symmetry and b
is odd exchange symmetry. The DCS plotted with the red mesh includes the geometric phase (GP)
while the one with the black mesh does not (NGP). The results include all values of total angular
momentum J = 0–4. Reproduced with permission from [44]
276 N. Balakrishnan and B. K. Kendrick
Acknowledgements N. B. acknowledges support from the Army Research Office, MURI grant
No. W911NF-12-1-0476 and the National Science Foundation, grant No. PHY-1505557. B. K. K.
acknowledges that part of this work was done under the auspices of the US Department of Energy
under Project No. 20140309ER of the Laboratory Directed Research and Development Program
at Los Alamos National Laboratory. Los Alamos National Laboratory is operated by Los Alamos
National Security, LLC, for the National Security Administration of the US Department of Energy
under contract DE-AC52-06NA25396.
References
36. Ospelkaus S, Ni K-K, Wang D, de Miranda MHG, Neyenhuis B, Quéméner G, Julienne PS,
Bohn JL, Jin DS, Ye J (2010) Science 327:853
37. Knoop S, Ferlaino F, Berninger M, Mark M, Nägerl H-C, Grimm R, D’Incao JP, Esry BD
(2010) Phys Rev Lett 104:053201
38. Balakrishnan N (2016) J Chem Phys 145:150901
39. Levinson N (1949) Kgl. Danske Videnskab Selskab Mat Fys Medd 25:9
40. Kendrick BK, Hazra J, Balakrishnan N (2015) Nat Commun 6:7918
41. Hazra J, Kendrick BK, Balakrishnan N (2015) J Phys Chem A 119:12291
42. Kendrick BK, Hazra J, Balakrishnan N (2015) Phys Rev Lett 115:153201
43. Hazra J, Kendrick BK, Balakrishnan N (2016) J Phys B At Mol Opt Phys 49:194004
44. Kendrick BK, Hazra J, Balakrishnan N (2016) New J Phys 18:123020
45. Kendrick BK, Hazra J, Balakrishnan N (2016) J Chem Phys 145:164303
46. Simbotin I, Ghosal S, Côté R (2011) Phys Chem Chem Phys 13:19148
47. Simbotin I, Ghosal S, Côté R (2014) Phys Rev A 89:040701
48. Simbotin I, Côté R (2015) N J Phys 17:065003
49. Pack RT, Parker GA (1987) J Chem Phys 87:3888
50. Kendrick BK, Pack RT, Walker RB, Hayes EF (1999) J Chem Phys 110:6673
51. Mead CA (1980) J Chem Phys 72:3839
52. Boothroyd AI, Keogh WJ, Martin PG, Peterson MR (1996) J Chem Phys 104:7139
53. Mielke SL, Garrett BC, Peterson KA (2002) J Chem Phys 116:4142
Part III
Biochemistry and Biophysics
Adducts of Arzanol with Explicit Water
Molecules: An Ab Initio and DFT Study
Liliana Mammino
L. Mammino (✉)
Department of Chemistry, University of Venda, P/Bag X5050, Thohoyandou 0950,
South Africa
e-mail: [email protected]
R" R'
O
H
1 Introduction
(comprising the prenyl chain attached to C5), PYR for the α-pyrone moiety, and
ARZ for arzanol.
The computational study of the ARZ molecule [5] showed that its conforma-
tional preferences are influenced by the patterns of intramolecular hydrogen bonds
(IHB), which are the dominant stabilising factor, by the mutual orientation of
the PHL and PYR moieties (which also determines part of the IHB patterns), by the
orientation of the phenol OHs (as is true for ACPLs in general [6–10]) and by the
orientation of the prenyl chain (which is generally true for prenylated ACPLs [8]).
The IHBs comprise the IHB formed by O14 and either H15 or H16 (here termed
“first IHB” [6–10]), the IHBs between the two moieties (which will be categorised
as IMHB, for ‘intermoiety H-bonds’, when it is relevant to underline this role, [5])
and the O10-H16⋯π or O12-H17⋯π interactions, when either O10-H16 or
O12-H17 and the prenyl chain have favourable orientations. The distribution of
donor and acceptor sites in the ARZ molecule enables the formation of two
simultaneous IMHBs, one on either side of the methylene bridge, and all the lowest
energy conformers are characterised by the presence of the first IHB and two
IMHBs [5]; when the first IHB engages H15, it is cooperative with the IMHB
engaging O8.
The computational study of ARZ [5] included calculations in three solvents
(chloroform, acetonitrile and water) utilising the Polarizable Continuum Model
(PCM, [11–13]). In general, “continuum solvation models are the ideal conceptual
framework to describe solvent effects within the QM approach” [13]. However,
PCM does not take into explicit account directional solute-solvent interactions such
as hydrogen bonding [14] (except implicitly for some effects [15]). On the other
hand, solute-solvent H-bonding is important for solute molecules containing
H-bond donors or acceptors and solvent molecules capable of forming H-bonds.
The most important of these solvents is water, which constitutes the highest pro-
portion of the mass of living organisms. The consideration of adducts with explicit
water molecules is the most informative option on solute-solvent H-bonding util-
ising QM approaches. It can provide information about preferential arrangements of
water molecules in the vicinity of the various donor or acceptor sites. It can also
contribute information on the outcome of the competition between intramolecular
H-bonding and intermolecular solute-water H-bonding through energetics com-
parisons (by comparing an adduct maintaining a certain IHB and an adduct in
which its donor or acceptor is engaged in a solute-solvent H-bond), and also
through the optimisation itself, which may ‘open’ (break) specific IHBs, as verified,
e.g., in the study of adducts of caespitate [16] or other ACPLs [17] with explicit
water molecules.
This work considers adducts of various conformers of ARZ with explicit water
molecules, trying to identify patterns for the energy of the solute-solvent interaction
at different binding sites and for preferred arrangements of water molecules around
different sites of the ARZ molecule. The study appears to be particularly interesting
because of the high number of H-bond donors and acceptors in the ARZ molecule
and because of the presence of IHBs (including cooperative ones), which influences
the way in which water molecules approach the corresponding regions. The work
284 L. Mammino
pertains to an ongoing study of antioxidant ACPLs which, for ARZ, has so far
produced the results reported in [5, 18].
2 Calculation Details
ARZ can have a high number of conformers, with different mutual orientations of
the PHL and PYR moieties and different IHB patterns. More than 90 conformers
were considered in [5] (which included all the lowest energy ones and disregarded
only some very high energy ones without the first IHB). Four mutual orientations of
the two moieties were identified and denoted with the numbers 1, 2, 3 and 4 [5].
Corresponding conformers of the #1 and #2 series differ only by the fact that the
orientation of the PYR ring is symmetrical with respect to the plane identified by
the PHL ring and have very close energies; the same is true for corresponding
conformers of the #3 and #4 series. Therefore, it is sufficient to select one series
from each pair to ensure the consideration of the different IHMB patterns. The #1
and #3 series were selected, to include the minimum energy conformers of each pair
(Fig. 3). It can be expected that adducts of conformers of the #2 series with the
same number and input-arrangement of water molecules as for corresponding
conformers of the #1 series will be very similar, as the PYR-related portion is
symmetrical in the two cases and the IHB and steric patterns are the same.
Equivalently, adducts of corresponding conformers of the #3 and #4 series are
expected to be very similar. A few adducts of conformers of the #2 and #4 series
were considered to verify this expectation.
1-d-r-ξ-αδ 3-s-w-η-γτ
Fig. 3 The two lowest energy conformers with different mutual orientations of the two ring
systems, as identified in [5], and the. Acronyms denoting them
Adducts of Arzanol with Explicit Water Molecules … 285
The variety of possible adducts with explicit water molecules is extremely high
for a solute molecule like ARZ, which contains a high number of H-bond donor or
acceptor sites and can form a variety of conformers in which one or another site
may be more or less available to form H-bonds with water molecules. It was opted
to consider all the lower energy conformers of the #1 and #3 series and repre-
sentative higher energy ones, to ensure that the most interesting arrangements of
water molecules around the various sites of ARZ are captured.
Two sets of adducts were considered. One set involves most of the conformers
of the #1 and #3 series, and entails adducts in which one water molecules is
attached in turn to each donor or acceptor site of each conformer. These adducts
enable a comparison of the energy with which each site can bind a water molecule,
and also offer indications about how close a water molecule can approach the given
site. In the real situation within the solvent, this is determined also by the inter-
actions between water molecules and, therefore, a water molecule attached to a
given site might remain at a greater distance than in the models with only one water
molecule. On the other hand, it may happen that one (or, sometimes, more) water
molecules remain attached to the solute molecule when it enters the active site of
the biological target and may contribute to the binding between the molecule and
the target; in such cases, the water molecule will likely remain as close as possible
to the ARZ site to which it binds. The knowledge of the strength with which a water
molecule binds to a certain donor or acceptor site of ARZ may thus be useful also
for a better understanding of its permanence (when it occurs) when ARZ binds to its
target, or its role in such binding.
The second set comprises adducts with several explicit water molecules,
attempting to approximate a first solvation layer or portions of it. Like in previous
studies on adducts of ACPLs with explicit water molecules [16, 17, 19], the ‘first
solvation layer’ concept is expanded to include not only the water molecules
directly H-bonded to suitable sites of the solute molecule, but also water molecules
that might bridge them (the presence of a third water molecule bridging two
molecules directly H-bonded to the solute often has a stabilizing effect [17]). The
distribution and spacing of the several H-bond donors or acceptors in ARZ enable
the possibility of considering adducts in which the water molecules attached to
ARZ, and those bridging them, approximate a continuous layer in the vicinity of
extensive portions of ARZ.
The inputs were prepared placing water molecules in the vicinity of H-bond
donor or acceptor sites of the selected conformers of ARZ. Different numbers and
arrangements of water molecules were considered for each conformer, also taking
into account the resulting arrangements of already optimised outputs. For instance,
when one or more water molecules ‘moved’ into a second solvation layer on
optimization (out of contact with the ARZ molecule and with no bridging role
between water molecules attached to it), those molecules were removed and the
resulting input was optimised as a new adduct.
The selection of the number/s of water molecule in the adducts likely to better
contribute the desired information is a rather delicate issue. Too small a number
would not enable the incorporation of the effects of water-water interactions
286 L. Mammino
relevant for the first solvation layer, such as the stabilising effect of a water
molecule bridging two water molecules H-bonded to ARZ. On the other hand, the
tendency of water molecules to cluster together limits the number of water mole-
cules in an adduct, if one wishes them to ‘remain’ in the first solvation layer on
optimisation. For the case of ARZ, it was found that when more than 9–10 water
molecules are present, their tendency to cluster becomes dominant and several of
them may move away from their initial binding sites of ARZ, yielding arrangements
in which they ‘crowd’ in the vicinity of only a portion of ARZ, and one or more of
them may move beyond the first solvation layer.
All the adducts were calculated in vacuo, performing optimisation with fully
relaxed geometry at the same levels utilised in all the previous calculations on
ACPLs [6–10, 16, 17, 19–25], i.e., Hartree Fock (HF) with the 6-31G(d,p) basis set
and Density Functional Theory (DFT) with the B3LYP functional [26–28] and the
6-31+G(d,p) basis set. The reasons for the selection of the two levels of theory and
basis sets are explained in the previous works [6–10, 16–24]; it is considered
important to maintain them in this and further studies involving ACPLs, to enable
informative and straightforward comparisons.
In the previous studies [5–10], DFT calculations have mostly been performed as
post-HF calculations; random testing had shown that the same inputs optimise to
the same conformers with the two methods; thus, treating DFT as post-HF calcu-
lations was expedient to decrease computational costs. In the case of the adducts of
ARZ considered here, HF and DFT calculations were performed independently for
each input because the presence of several H-bond donor and acceptor sites and the
non-covalent nature of the solute-solvent H-bonds suggests the possibility that the
two methods may lead to different arrangements of water molecules. In most cases,
the same input optimised to similar arrangements, but, in a number of cases, the
optimised adducts differed substantially. Such outputs were then utilised as inputs
for the other method, what enabled the consideration of additional geometries that
had not been envisaged on the initial input-preparation.
The interaction energy (ΔEarz-n ⋅ aq) between the ARZ molecule and the water
molecules bonded to it was calculated for each adduct. The general equation is [29]
ΔEarz−n ⋅ aq = Eadduct − Earz + n Eaq − ΔEaq−aq ð1Þ
where Eadduct is the energy of the adduct, n is the number of water molecules in the
adduct, Earz is the energy of the isolated ARZ conformer, Eaq is the energy of an
isolated water molecule and ΔEaq-aq is the overall interaction energy between water
molecules, resulting mainly from water-water H-bonds.
Adducts of Arzanol with Explicit Water Molecules … 287
Fig. 4 Example of an adduct of arzanol with 9 explicit water molecules and the system of the sole
water molecules in the same arrangement as in the adduct, used to calculate ΔEaq-aq
For adducts with only one water molecule, the equation becomes simply
For adducts with interacting water molecules (which is the most common case
when there are several water molecules), ΔEaq-aq is evaluated through a single point
(SP) calculation on a group of water molecules arranged exactly as in the adduct,
but without the ARZ molecule ([29], Fig. 4). If Eaq-set is the energy of this set of
water molecules, then
Both Eadduct and Eaq-set have been corrected for basis set superposition error
(BSSE), using the counterpoise method [30], when the adduct contains more than
one water molecule. The BSSE correction was not applied to adducts containing
only one water molecule because it would be small for these adducts and because
these calculations are meant for comparisons. The values for the calculated adducts
with four or more water molecules show that the BSSE correction increases sub-
stantially as the number of water molecules in the adduct increases; therefore, it will
be smallest for adducts with one water molecule (also in view of the absence of
288 L. Mammino
water-water interactions). The main objective of calculating these adducts was that
of enabling comparison of the binding strengths of individual sites. Neglecting
BSSE corrections does not appear to greatly affect comparisons among analogous
molecular systems when the correction is sufficiently small [31] (all the adducts
with one water molecule consist of one ARZ molecule and one water molecule, so,
they are analogous molecular systems). It was thus assumed that the comparisons
among these adducts remain reliable also without BSSE correction. Neglecting the
correction proved expedient also in view of the high number of calculated adducts
of this type.
All the calculations were performed with GAUSSIAN 03, Revision D 01 [32].
All the energy values reported are in kcal/mol and all the distances are in Å. For
conciseness sake, the calculation methods will be denoted simply as HF for HF/
6-31G(d,p) and DFT for DFT/B3LYP/6-31+G(d,p) in the text.
Detailed information, including tables with all energy values and H-bond
parameters for the calculated adducts, and figures showing the geometries of the
calculated adducts, is provided in the Supplementary Information.
3 Results
Following a practice introduced since the initial studies of ACPLs [6–10], the
conformers are denoted by acronyms which provide information about their
characteristics, to enable easy tracking of geometric characteristics and energy-
influencing features. The acronyms for the ARZ conformers start with the number
denoting the mutual orientation of the moieties. The other geometry features are
denoted by letters, whose meanings are listed in Table 1 [5]. The two lowest energy
conformers and their acronyms are shown in Fig. 3.
In order to identify them in a straightforward way, the adducts are denoted by the
acronym of the ARZ conformer at their centre, followed by information about the
water molecules. For the adducts with one water molecule, the name of the con-
former is followed by ‘1aq’ and by the position to which the water molecule is
attached. Thus, 1-d-r-ξ-αδ-1aq-O14 denotes the adduct of the 1-d-r-ξ-αδ conformer
in which one water molecule is attached to O14; 1-d-r-αδ-1aq-H17-π1 denotes the
adduct of 1-d-r-αδ in which a water molecule is attached to H17 and is also
interacting with the π bond of the prenyl chain; and so on. The adducts containing
more than one water molecule are denoted with the name of the conformer followed
by the number of molecules in the adduct. For instance, 1-d-r-ξ-αδ-5aq denotes an
adduct of conformer 1-d-r-ξ-αδ with 5 water molecules. Since different arrange-
ments of the water molecules are possible for adducts of the same conformer and
with the same number of water molecules, letters are added at the end of the
acronym to distinguish them from one another. However, the information on the
site to which each water molecule is attached would be too bulky to summarise it in
an acronym, and 3D models showing the arrangements of the water molecules are
Adducts of Arzanol with Explicit Water Molecules … 289
Table 1 Symbols utilised to specify geometrical characteristics in the acronyms denoting the
conformers. The symbols d, s, r, w, u, η and ξ have the same meanings as in other studies on
ACPLs [6–10] and the symbols α, β, γ, δ, ε and τ had been introduced in [5]
Symbol Meaning
d The H15⋯O14 first IHB is present
s The H17⋯O14 first IHB is present
r H16 is oriented towards the α-pyrone ring
w H16 is oriented towards the prenyl chain
u H15 or H17, not engaged in the first IHB, is oriented toward the COR group
(‘upwards’)
η Presence of O-H⋯π interaction between H16 and the C29=C30 double bond
ξ Presence of O-H⋯π interaction between H17 and the C29=C30 double bond
a No O-H⋯π interaction is present, and the prenyl chain is oriented ‘upwards’
b No O-H⋯π interaction is present, and the prenyl chain is oriented ‘downwards’
α The H27⋯O8 intermonomer hydrogen bond is present
β The H15⋯O26 intermonomer hydrogen bond is present
γ The H15⋯O23 intermonomer hydrogen bond is present
δ The H16⋯O23 intermonomer hydrogen bond is present
ε The H16⋯O26 intermonomer hydrogen bond is present
τ The H27⋯O10 intermonomer hydrogen bond is present
π1 The C29=C30 double bond, when acting as binding site for a water molecule
the only option to convey a clear image of each adduct (the 3D models are included
in the Supplementary Information).
Given the high number of conformers of the ARZ molecule, and the high number of
sites to which a water molecule can bind (including the possibility of simultaneous
binding to two geometrically suitable sites), considering all the possible adducts of
this type would be unaffordable. It was opted to calculate a selection of sufficiently
representative adducts for each binding site. While some sites (e.g., O14 or O23)
are accessible in all conformers, other sites are not available in some conformers,
and the combinations for simultaneous binding to two sites depend on the type of
conformer; therefore, only a limited number of adducts may be obtainable for
certain combinations. Changes during optimisation reduce the number of adducts
for some sites while increasing it for others (which informs that the former sites or
site-combinations are less favourable than others).
A total of more than 200 adducts were calculated. Figure 5 shows the main
geometries obtained, according to the type of conformer of the ARZ molecule and
to the site to which the water molecule binds. All the geometries are shown in the
290 L. Mammino
Fig. 5 Representative adducts of arzanol with one water molecule attached to different donor or
acceptor sites. The first two rows show adducts of the lowest energy conformer of arzanol
(1-d-r-ξ-αδ), the other rows show other relevant types of arrangement of the water molecule,
for different types of conformers
Supplementary Information. Nearly all the geometries are obtained with both HF
and DFT, with similar shapes.
In many cases, the optimisation does not change the site to which the water
molecule binds; it is the case of inputs in which it binds to O14 or to O23, but also
to other favourable sites. There are, however, also several cases in which the
Adducts of Arzanol with Explicit Water Molecules … 291
Table 2 Examples of changes occurring during optimization, highlighting the binding prefer-
ences of the water molecule
Reference Input geometry Output geometry
acronym
ch-1 1-d-r-ξ-αδ-1aq-O19-O23 1-d-r-ξ-αδ-1aq-O23
ch-2 1-d-r-ξ-αδ-1aq-O8 1-d-r-ξ-αδ-1aq-O14
ch-3 3-s-r-b-γε-1aq-O12 3-s-r-b-γε-1aq-O14
ch-4 3-s-w-η-γτ-1aq-O8 3-s-w-η-γτ-1aq-O23
ch-5 1-d-r-b-αδ-1aq-O8 1-d-r-b-αδ-1aq-O26
ch-6 3-d-w-η-τ-1aq-O8 3-d-w-η-τ-1aq-O8-O23
ch-7 3-d-w-ξ-1aq-O23 3-d-w-ξ-1aq-O8-O23
ch-8 3-d-r-ξ-ε-1aq-O26 3-d-r-ξ-ε-1aq-H27
ch-9 3-d-w-η-τ-1aq-O12 3-d-w-η-τ-1aq-H17
ch-10 3-d-w-ξ-τ-1aq-O10 3-d-w-ξ-τ-1aq-H16
ch-11 1-s-r-βδ-1aq-O8 1-s-r-βδ-1aq-O8-H27
ch-12 1-s-r-βδ-1aq-O26 1-s-r-βδ-1aq-O8-H27
ch-13 3-d-w-η-τ-1aq-O26 3-d-w-η-τ-1aq-O26-H16
ch-14 3-d-r-u-b-ε-1aq-O26 3-d-r-u-b-ε-1aq-O10-H27
ch-15 1-s-r-u-δ-1aq-O8-O26 1-s-r-u-βδ-1aq-O8-H27
ch-16 3-s-r-γε-1aq-O10-π1 3-s-r-γε-1aq-O10-H27
ch-17 3-s-w-η-γ-1aq-O10-O26 3-s-w-η-γ-1aq-O10-H27
ch-18 1-d-w-ξ-α-1aq-O10-O23 1-d-ξ-α-1aq-H16-O23
ch-19 1-d-w-η-α-1aq-O10-O23 1-d-w-η-α-1aq-H16-O23
ch-20 1-s-w-u-η-α-1aq-H15 1-s-w-u-η-α-1aq-H15-O26
ch-21 3-s-w-u-η-τ-1aq-H15 3-s-w-u-η-τ-1aq-H15-O23
ch-22 1-d-w-α-1aq-H16 1-d-w-α-1aq-H16-O23
ch-23 1-s-r-βδ-1aq-H27 1-s-r-βδ-1aq-O8-H27
ch-24 1-d-w-ξ-1aq-H27 1-d-w-ξ-1aq-O14-H27
ch-25 3-s-w-η-γ-1aq-H27 3-s-w-η-γ-1aq-O10-H27
ch-26 1-d-w-ξ-α-1aq-O12 1-d-w-α-1aq-O12
ch-27 1-d-r-b-αδ-1aq-O10-π1 1-d-r-ξ-αδ-1aq-O10’
ch-28 1-d-w-η-α-1aq-O10-O23 1-d-w-ξ-α-1aq-O12
ch-29 1-d-r-b-αδ-1aq-H17 1-d-r-αδ-1aq-H17-π1
ch-30 3-s-w-a-γτ-1aq-H16 3-s-w-γτ-1aq-H16-π1
ch-31 3-d-w-τ-1aq-H16-π1 3-d-w-ξ-τ-1aq-H16
ch-32 1-d-w-1aq-H17-π1 3-d-w-η-τ-1aq-H17’
ch-33 3-s-w-a-γτ-1aq-O12-π1 3-s-w-η-γτ-1aq-O12
ch-34 3-d-r-b-ε-1aq-O10-π1 3-d-r-ξ-ε-1aq-O10-H27
optimisation changes the site to which the water molecule binds, with respect to the
input. Since these changes highlight binding preferences, they are given detailed
attention in the current analysis. Examples are reported in Table 2, considering
types of changes which appeared more than once. The changes are denoted with
292 L. Mammino
acronyms (‘ch’, which stands for ‘change’, followed by a number) to enable easy
referencing to them within this text.
The water molecule appears to prefer to bind to an sp2 O rather than to an sp3 O.
Thus, inputs in which the water molecule is attached to O19, or to both O19 and
O23, optimize to adducts with the water molecule attached to O23 (ch-1). Inputs in
which the water molecule is attached to the donor of the first IHB (O8 or O12, for d
and s conformers respectively), in conformers in which O14 is the only sp2 O in the
vicinity, often optimize to adducts in which water binds to O14 (ch-2, ch-3). If O23
is also available in the vicinity (e.g., in inputs involving 3-d conformers), the water
molecule often shifts from O8 to O23 (ch-4) or, sometimes, to O26 (ch-5). These
changes may be related to the hydrophobic character of IHB regions for hydrox-
ybenzenes in general [33], and of the first IHB of ACPLs in particular [17].
A clear preference appears for the water molecule to bind to two sites simul-
taneously, when two sites are geometrically suitable; this may lead to adducts in
which it binds to an sp2 O and an sp3 O simultaneously (e.g., O8 and O23; ch-6,
ch-7). The water molecule also appears to prefer to be acceptor to an OH group
(consistently also with the known tendency of phenol OHs to be donors in inter-
molecular H-bonds [16, 31]). This may involve rotation of H15, H16, H17 or H27,
to enable the formation of the H-bond with the water molecule (it is interesting to
recall that the phenol OHs in ACPLs do not usually rotate to form IHBs [16],
whereas the adducts calculated here show that the OHs in ARZ may rotate to form
an intermolecular H-bond with the solvent). The water molecule may change
binding site completely in order to be acceptor to the H of an OH (ch-8, ch-9,
ch-10). On the other hand, its tendency to bind to two geometrically suitable sites
simultaneously may lead to adducts in which it binds simultaneously to the H of an
OH and to a suitably close O atom (ch-11 to ch-19). The tendency to bind to two
sites appears also for inputs in which the water molecule is initially placed as
acceptor to an OH (ch-20 to ch-25).
The water molecule does not break O-H⋯O IHBs. It appears, however, to be
able to act on the O-H⋯π interaction, by breaking it (ch-26) or prompting it
(ch-27), or changing its pattern (ch-28, in which it changes from η to ξ). The water
molecule itself may interact with the C29=C30 π bond; the interaction usually
appears during optimisation (ch-29, ch-30), whereas inputs having the interaction
do not often optimise to adducts in which it is maintained (ch-31 to ch-34).
The changes just outlined appear both with HF and with DFT optimization. In
most cases, the same change occurs with the same input, i.e., HF and DFT opti-
misations lead to the same changed output. In some cases, the changes are different
with the two methods, leading to different outputs. Figure 6 shows a case in which
HF and DFT lead to outputs that are different from the input and different from each
other. In such cases, it is not easy to estimate a priori which of the two outcomes
might be closer to reality; for the specific case shown in Fig. 6, the HF result seems
more probable, as it shows preference for an sp2 O, which can form a stronger
H-bond than an sp3 O like O8 or O26.
Table 3 reports the relative energy and the ARZ-water interaction energy for
representative adducts, selected in such a way as to comprise nearly all the
Adducts of Arzanol with Explicit Water Molecules … 293
HF DFT
Fig. 6 An example in which the HF and DFT optimisations of the same input with one water
molecule lead to different results. In the input, the water molecule was attached to O8. The HF
optimisation moves it to O14 and the DFT optimisation moves it to O26. The conformer is the
lowest energy conformer of arzanol (1-d-r-ξ-αδ)
identified binding options of the water molecule. The relative energy depends to a
considerable extent on the type of conformer and its relative energy in vacuo,
whereas the interaction energy depends largely on the binding site (although the
geometry of the conformer influences the approach of the water molecule to a given
binding site). Table 4 reports the ranges of the interaction energy for the different
binding sites. The interaction energy values confirm the binding preferences of the
water molecule highlighted by the changes occurring during optimisation, such as
the preference for simultaneous binding to two atoms of ARZ (and, among these,
for one of the sites being a donor OH), and the preference for H atoms of OH
groups, followed by sp2 O atoms, when it binds only to one site.
Table 5 reports the ranges of the length of the ARZ-water H-bonds. The binding
sites are listed in the same sequence as in Table 4 to facilitate comparison in terms
of length of the H-bond and strength of the interaction energy (a few sites present in
Table 4 are not reported in Table 5 for space reasons). The correspondence is
meaningful because, for adducts with only one water molecule, the molecule-water
interaction energy can be viewed as the energy of the molecule-water H-bond, and
an H-bond length is an indication of its strength. The H-bond lengths are shorter
when the water molecule is acceptor to an OH group of ARZ. The H-bond lengths
for sp2 O are shorter than those of sp3 O, consistently with the ability of sp2 O to
form stronger H-bonds. Although the trends (comparison of lengths across adducts)
are largely similar in the HF and DFT results, the HF values are longer than the
DFT values for the same adducts. This is consistent with the known tendency of HF
to underestimate the strength of H-bonds and of DFT to overestimate it; therefore, it
appears reasonable to assume that the actual H-bond distance for a given adduct
will be somehow intermediate between the HF and the DFT values.
Table 3 Relative energy and arzanol-water interaction energy of representative adducts of arzanol with one explicit water molecule, in the HF/6-31G(d,p) and
294
Table 4 Ranges of the magnitude (absolute values) of the arzanol-water interaction energy for the
adducts of arzanol with one explicit water molecule, in the HF/6-31G(d,p) and DFT/B3LYP/6-31
+G(d,p) results in vacuo. When only two values are available, they are reported individually,
separated by a comma. When only one value is available, it is reported individually
Binding HF range DFT range Binding HF range DFT range
site (kcal/mol) (kcal/mol) site (kcal/mol) (kcal/mol)
O23-H15 15.4, 17.1 15.2, 16.8 H16-π1 7.2–8.7 7.3–9.9
O10-H27 7.6–15.6 9.3–12.1 O12 3.2–8.1 3.0–8.5
O23-H16 10.2–15.3 9.2–16.3 O8-O23 6.5–7.9 7.2–7.8
O8-H27 10.1–13.5 9.9 O23 3.8–7.8 4.0–11.1
O8-O26 12.3 5.4, 5.4 O14-O23 7.8 7.0
H17-π1 7.3–11.9 7.0–8.8 O12-O14 7.2, 7.5 7.5–7.8
O10-π1 6.1, 11.6 O14 3.9–7.1 3.7–7.4
O10-O26 11.6 6.1, 6.2 O8-O14 7.1 7.0
O26-H15 9.5–10.5 6.1–8.9 O10 3.9–7.0 5.2–8.2
H15 8.4–10.2 7.2 O26-H16 6.9
O10-O23 10.1 7.0–9.9 O12-π1 5.1–5.3 5.1–7.5
H27 5.7–10.0 6.6–11.2 O26 2.4–5.2 3.5–8.0
H16 8.3–9.5 7.6–9.9 O12-H17 4.6
O14-H27 8.7, 9.2 19.1 O8 4.3–4.4 4.5–5.1
H17 6.7–9.1 6.7–7.5 O14-O26 3.7
3.3 Results for the Adducts of Arzanol with More Than One
Water Molecule
A total of 84 adducts of ARZ with 4–12 water molecules were calculated, with
greater number of adducts with 7, 8 or 9 water molecules. What the calculation of
such adducts may contribute in a study of this type is not an exhaustive inclusion of
all possible options, but a sufficiently representative selection highlighting the
variety of possibilities and some recurrent patterns. The geometries present a high
variety of possible arrangements of the water molecules around the ARZ molecule.
Figure 7 shows some geometries of adducts with different numbers of water
molecules.
Figure 8 illustrates some relevant aspects in the arrangement of water molecules,
including some recurrent patterns. The water molecules tend to keep away from the
region of an IHB, as it is a hydrophobic region [17, 33]. Shapes already encoun-
tered for adducts of other ACPLs [17] appear with a certain frequency, such as a
pentagon of O atoms formed by the atoms engaged in an IHB and the O atoms of
the water molecules keeping away from it. The whole region of two cooperative
IHBs appears to be hydrophobic, and the water molecules may form a larger ring
(e.g., seven O atoms) while keeping away from it.
Differently from the adducts with one water molecule, a water molecule may
break an O-H⋯O IMHB and insert itself between the donor and the acceptor; the
Adducts of Arzanol with Explicit Water Molecules … 297
Table 5 Ranges of the length of the hydrogen bonds between the arzanol molecule and the water
molecule for selected binding sites in the calculated adducts of arzanol with one water molecule.
When only two values are available, they are reported individually, separated by a comma. When
only one value is available, it is reported individually
Binding site/s Length considered Range of values (Å)
HF results DFT results
O23, H15 Oaq⋯H15 1.878, 1.892 1.713, 1.754
Haq⋯O23 1.918, 1.953 1.747, 1.819
O10, H27 Oaq⋯H27 1.913–2.229 1.784–2.174
Haq⋯O10 1.983–2.298 1.911–2.282
O23, H16 Oaq⋯H16 1.851–2.190 1.690–1.751
Haq⋯O23 1.904–2.164 1.731–1.984
O8, H27 Haq⋯O8 1.966–2.125 –
Oaq⋯H27 1.877–1.932 –
O8, O26 Haq⋯O8 – 2.177–2.401
Haq⋯O26 – 2.033–2.177
H17, π1 H17⋯ Oaq 1.885–1.901 1.792–1.848
O26, H15 Oaq⋯H15 1.943–1.991 1.871–1.930
Haq⋯O26 2.365–2.429 2.035–2.199
H15 Oaq⋯H15 1.909–2.030 1.888–1.895
H27 Oaq⋯H27 1.890–2.009 1.822–1.933
H16 Oaq⋯H16 1.926, 1.935 1.905, 1.914
O14, H27 Haq⋯O14 2.253, 2.332 –
Oaq⋯H27 2.011, 2.025 –
H17 Oaq⋯H17 1.939–2.005 1.873–1.941
H16, π1 Haq⋯O10 1.879–1.901 1.752–1.870
O12 Haq⋯O12 2.132–2.240 1.931–2.060
O8, O23 H′aq⋯O23 2.120–2.158 1.965–2.037
Haq⋯O8 2.311–2.378 2.130–2.351
O23 Haq⋯O23 2.038–2.178 1.911–2.024
O14, O23 H′aq⋯O23 2.129, 2.182 2.008
Haq⋯O14 2.573, 2.604 2.662
O12, O14 H′aq⋯O14 2.217, 2.289 2.438–2.570
Haq⋯O12 2.285, 2.309 2.000–2.056
O14 Haq⋯O14 2.040–2.089 1.870–1.929
O8, O14 H′aq⋯O14 2.209 2.097
Haq⋯O8 2.366 2.220
O10 Haq⋯O10 2.106–2.158 1.952–2.023
O12, π1 Haq⋯O12 2.366–2.478 1.837–2.145
O26 Haq⋯O26 2.152–2.257 1.987–2.052
O8 Haq⋯O8 2.087–2.112 1.926–1.957
298 L. Mammino
Fig. 7 Illustration of the variety of possible geometries for the arrangement of water molecules in
the adducts
Adducts of Arzanol with Explicit Water Molecules … 299
Fig. 8 Some relevant aspects in the arrangement of water molecules around the arzanol molecule.
The water molecules tend to keep away from the IHB regions; a frequent result is a pentagon of O
atoms (including those of the IHB), as in 2-s-r-βδ-9-aq-x, 4-s-r-a-γε-8aq-x and 3-s-w-a-γτ-9aq; in
some cases, the ring may contain more O atoms, as the ring around two consecutive IHBs in
2-d-w-η-α-6aq. A water molecule may break an inter-monomer H-bond, as in 3-s-w-a-γτ-9aq,
where it breaks the H27⋯O10 IHB. In a number of cases, the tendency of the water molecules to
cluster together fosters arrangements in which some donor or acceptor sites of the arzanol
molecule are not binding a water molecule, as in 3-s-w-b-γτ-9aq-e (where no water molecule
attaches to O14, although it is a strong acceptor) and in 3-s-w-b-γτ-11aq. Although in adducts with
one or few water molecules no water molecule binds to O19, it may happen that a water molecule
binds to it if there are enough bridging water molecules to facilitate the arrangement, as in
3-s-w-b-γτ-11aq
cases observed concern IMHBs where H27 is the donor (H27⋯O8 or H27⋯O10).
The length of the H-bond between H27 and the O of the water molecule attached to
it is the shortest ARZ-water H-bond length in these adducts, suggesting that this
H-bond might be among the strongest if not the strongest. No breaking has been
observed for the first IHB (which is considerably stronger than H27⋯O8 or
H27⋯O10), consistently with the results for the adducts of other ACPLs [16, 17].
For similar reasons (the acceptor being an sp2 O), no breaking is observed for
IMHBs in which O23 is the acceptor. The O-H⋯π interaction is broken in a number
of cases, with a water molecule inserting itself between the H and the C29=C30
π-bond. The tendency of the water molecules to cluster together may prompt major
changes during optimization, leading to chains of water molecules alternatingly
binding to a site of ARZ and playing bridging roles.
The relative energies of adducts with the same number of water molecules
depend on their geometrical arrangement and binding sites more than on the relative
energy of the isolated conformer. The BSSE correction increases as the number of
300 L. Mammino
Table 6 Ranges of the BSSE correction to the energy of the adduct and to the energy of the
system of the sole water molecules, and percentage of the latter correction with respect to the
former, for the calculate adducts of arzanol with explicit water molecules. The ranges are
considered according to the number of water molecules in the adduct
Number of BSSE correction BSSE correction range for Range of percentage
water range for the adduct the sole water molecules of the contribution
molecules (kcal/mol) (kcal/mol) of water
HF DFT HF DFT HF DFT
5 8.50–11.05 5.15–6.46 0.95–3.99 0.99–1.84 9.63–47.25 16.56–55.52
6 10.32–12.84 6.14–8.11 2.25–6.64 1.61–3.51 17.84–53.74 25.17–50.87
7 12.52–15.45 7.34–9.71 3.00–7.70 2.65–5.83 21.00–52.65 31.93–62.99
8 12.45–18.28 8.43–11.17 4.85–8.91 4.17–6.01 30.57–57.59 45.51–71.32
9 14.62–21.96 8.94–13.06 5.53–11.12 4.48–7.99 36.58–57.60 50.11–64.28
10 19.87–20.42 11.16–14.52 11.86–12.29 7.25–9.39 50.84–61.86 60.18–64.97
11 20.00–25.85 14.68–15.70 9.23–13.97 9.37–10.52 45.41–55.07 53.68–67.01
12 24.71–25.77 15.81–16.51 11.46–12.62 8.42–9.75 46.31–49.18 53.30–60.65
water molecules in the adduct increases, although it can vary broadly for adducts
with the same number of water molecules. The ranges of values are somewhat
narrower in the DFT results and somewhat broader in the HF results (Table 6). The
correction is often greater for adducts with lower relative energy.
The ARZ-water interaction energy depends on the conformer of ARZ, on the
sites to which the water molecules bind and on the number of water molecules. The
dominant factor appears to be the arrangement of the water molecules, which
includes the binding sites and also the types of water-water interactions. Table 7
reports the ranges of the ARZ-water interaction energy as identified from the cal-
culated adducts; the values show a tendency to higher upper-limit of the range as
the number of water molecules increases, but without a straightforward relationship
between the two. The dominance of qualitative aspects (characteristics of the ARZ
conformer, types of binding sites, geometry of the water molecules arrangements)
may hamper the possibility of identifying more definite types of relationship. The
magnitude of the DFT values tend to be greater than that of the HF values for
corresponding conformers, which may be ascribed to the tendency of HF and DFT
to respectively underestimate and overestimate the strength of H-bonds.
The length of the ARZ-water H-bonds depends on the binding site and appears to
be fairly consistent with the patterns highlighted by the adducts with one water
molecule, although the other water molecules binding to a water molecule attached to
ARZ may influence the length of its H-bond with the ARZ site. This is the case of
H-bonds between H27 and the O of a water molecule, which are the shortest in
adducts with several water molecules, but not in the adducts with one water molecule.
Adducts of Arzanol with Explicit Water Molecules … 301
Table 7 Ranges of the magnitude of the arzanol-water interaction energy (corrected for BSSE)
according to the number of water molecules in the adduct
Number Magnitude of the Number Magnitude of the
of water interaction energy of water interaction energy
molecules (kcal/mol) molecules (kcal/mol)
HF DFT HF DFT
5 13.64–30.52 15.82–36.70 9 16.14–38.63 21.10–45.31
6 13.40–35.77 17.93–39.47 10 16.79–24.24 18.92–33.62
7 6.32–35.95 8.64–48.69 11 22.16–37.55 24.23–47.61
8 11.97–41.16 15.16–41.93 12 26.13–39.39 31.12–57.73
A study of the type considered in this work involves a variety of challenges because
of the nature of the adducts and of the character of the information that can be
obtained.
The high number of conformers of the ARZ molecule and the high number of its
H-bond donor or acceptor site implies an enormous number of possible adducts
with a given number of water molecules, and this number increases rapidly as the
number of water molecules in the adducts increases. Calculations show high sen-
sitivity of the optimisation procedure to small differences in the inputs, which
would recommend the consideration of several adducts with similar or very similar
(but not identical) input geometries for each relevant geometry, thus further mul-
tiplying the number of potentially interesting adducts. On the other hand, the ten-
dency of water molecules to cluster on optimisation may yield similar adducts from
different inputs, which decreases the informative role of the result. It may also lead
to adducts where some water molecules cluster beyond the boundaries of the
adopted criterion for the definition of ‘first solvation layer’, resulting in high
water-water interaction energy and poor solute-water interaction energy; since the
latter phenomenon is more extensive as the number of water molecules in the
adduct increases, it prevents the possibility of considering a higher number of water
molecules than the one for which their clustering becomes extensive or dominant.
The huge number of adducts that would be needed to provide a comprehensive
panoramic taking into account all the relevant energy-influencing features (all the
geometrical features of all the conformers of ARZ, and all the possible arrange-
ments of water molecules around each conformer) would imply enormous com-
putational costs. Therefore, it was opted to select representative adducts for each
relevant characteristic. This corresponds to a sampling approach more than to an
unaffordable exhaustive approach. All the same, a sampling approach can be
informative for a variety of aspects.
Within the reality of a water solution, the supermolecular structures of the
adducts are not ‘fixed’ in time, because the water molecules H-bonded to a solute
molecule do not remain the same in time (there is continuous fast interchange with
the surrounding water molecules). Therefore, the calculated adducts represent
time-averaged probable possibilities rather than permanent structures.
Despite all these challenges, the calculation of adducts with explicit water
molecules provides information on the relative strength with which a water mole-
cule can bind to each donor or acceptor site of the solute molecule, on the distance
to which a water molecule preferably approaches each site, and on the preferred
arrangements of water molecules around each donor or acceptor sites or around the
region of two or more spatially close donors or acceptors. Since it results from
optimisation and H-bonds are directional, this information can be considered as
responding to the more common situations in the vicinity of the donors or acceptors
of the solute molecule. The adducts also provide indications about whether a certain
IHB tends to remain or to break in water solution—a type of information which
Adducts of Arzanol with Explicit Water Molecules … 303
References
32. Frisch MJ, Trucks GW, Schlegel HB, Scuseria GE, Robb MA, Cheeseman JR, Mont-
gomery JA, Vreven T, Kudin KN, Burant JC, Millam JM, Iyengar SS, Tomasi J, Barone V,
Mennucci B, Cossi M, Scalmani G, Rega N, Petersson GA, Nakatsuji H, Hada M, Ehara M,
Toyota K, Fukuda R, Hasegawa J, Ishida M, Nakajima T, Honda Y, Kitao O, Nakai H,
Klene M, Li X, Knox JE, Hratchian HP, Cross JB, Adamo C, Jaramillo J, Gomperts R,
Stratmann RE, Yazyev O, Austin AJ, Cammi R, Pomelli C, Ochterski JW, Ayala PY,
Morokuma K, Voth GA, Salvador P, Dannenberg JJ, Zakrzewski VG, Dapprich S,
Daniels AD, Strain MC, Farkas O, Malick DK, Rabuck AD, Raghavachari K, Foresman JB,
Ortiz JV, Cui Q, Baboul AG, Clifford S, Cioslowski J, Stefanov BB, Liu G, Liashenko A,
Piskorz P, Komaromi I, Martin RL, Fox DJ, Keith T, Al-Laham MA, Peng CY,
Nanayakkara A, Challacombe M, Gill PMW, Johnson B, Chen W, Wong MW, Gonzalez C,
Pople JA (2003) Gaussian 03. Gaussian Inc, Pittsburgh
33. Mammino L, Kabanda MM (2011) Int J Quant Chem 111:3701–3716
Computational Study of Jozimine A2,
a Naphthylisoquinoline Alkaloid
with Antimalarial Activity
Keywords Alkaloids ⋅
Antimalarials ⋅
Intramolecular hydrogen bond
Jozimine A2 ⋅ Naphthyl-isoquinoline alkaloids ⋅
Mutual orientation
Solute-solvent interactions
1 Introduction
57
CH3
58 35
H 36
N 34
A'
37 39
H359C
40 33
B'
60
H 38 32
46
O 31
56
24 23 CH3
25 29 22 52 51
H CH3
C' D' 43 42
O O
26 30 21
27 28 14 15
13 19 16
O45 O44 D C
H355C 54
H 12 20 17
H353C 11 18
41
5 O
6 4 H50
B
7 9 49
CH3
10 3
A
2
8 N
1 H48
47
CH3
Fig. 1 Two representations of the jozimine A2 molecule and atom numbering utilized in this
work. In the structure on the left, the C atoms in the rings are represented only by the numbers
denoting their positions. Only the H atoms attached to O or N are numbered separately, while the
H atoms attached to C atoms are given the same number as the C atom and are not shown in the
structure. The two units are identical. Identical rings are denoted by the same letters, primed for
one of the units (A, B, C, D and C′, D′, A′ and B′)
were used to perform the calculations. Calculations in solution were performed using
the PCM model.
The results highlight the dominant stabilizing effect of the O−H⋯O
intramolecular hydrogen bonds (IHBs) and of other H-bond-type intramolecular
interactions (for instance, O−H⋯π interaction) and the preference for all the
moieties to be mutually perpendicular. Other computable molecular properties
(dipole moments, HOMO-LUMO energy gaps, solvent effect, etc.) also show some
dependence on IHB patterns.
Comparisons are also made with the results of a study of michellamine A [12]—
a dimeric naphthylisoquinoline alkaloid with anti-HIV activity [13] and with the
same ring systems as JZM. The two molecules differ by the substituents in the
moieties and by the fact that the two naphthalene moieties in michellamine A are
not identical. The stabilising effects show interesting similarities. The main dif-
ference is that no IHB between the two naphthalene moieties is possible in JZM,
whereas it is possible in michellamine A.
308 M. K. Bilonda and L. Mammino
2 Computational Details
Calculations in vacuo were performed with fully relaxed geometry using two levels
of theory: Hartree-Fock (HF) with the 6-31G(d,p) basis set, and Density Functional
Theory (DFT) with the B3LYP functional [14, 15] and the 6-31+G(d,p) basis set.
HF is a moderately cheap quantum mechanical method which can yield accurate
information regarding conformational analysis. Previous studies on other molecules
[12, 16–18] showed that HF can successfully handle intramolecular H-bonding and
yields HOMO-LUMO energy gaps approaching those of experiments. DFT—an
alternate method to wavefunction approaches—is often used in conformational
search because it takes into account part of the correlation effects at a relatively low
cost. Among the numerous functionals available for the DFT framework, B3LYP
[14, 15, 19] is the most widely utilized; it can provide better quality results in
combination with basis sets containing diffuse functions, above all for molecular
systems containing IHBs [12, 16–18].
Harmonic vibrational frequencies were calculated in vacuo at the HF/6-31G(d,p)
level to verify that the stationary points from optimization results corresponded to
true minima and to obtain the zero-point energy (ZPE) corrections. The frequency
values were scaled by 0.9024 [20].
A preliminary identification of conformers of interest was carried out by con-
sidering the potential energy profiles for the rotation of the C5−C11 bond (showing
the minima for the mutual orientation of the two moieties within one unit, Fig. 2)
and for the rotation of the C13−C21 bond (showing the minima for the mutual
orientation of the two units). The rotation of the two bonds was carried out
simultaneously, yielding a 3D potential energy profile (Fig. 2). It was carried out
17.50
315
80 31.25
tive e
45.00
270 58.75
n
60
erg y
72.50
Torsion angle D1
225 86.25
40 100.0
(kcal/
180
20
m
ol)
0 360 135
0 315
45 270 90
D2
90 225
gle
135 180 45
To
an
rs 180 135
io n
io n
225
an 90 0
rs
45
D1 315
360 0 Torsion angle D2
Fig. 2 3D potential energy profile for the scan of the rotation of the C5–C11 and C13–C21 bonds
(energy versus D1 and D2, with D1 being the C4–C5–C11–C12 torsion angle and D2 being the
C14–C13–C21–C28 torsion angle) and 2D potential energy profile for the same rotation scan. The
complete 360° rotations were performed at 15° pace
Computational Study of Jozimine A2, a Naphthylisoquinoline … 309
for a complete rotation (360°) of the bond with a 15° pace. A conformer without
O−H⋯π interaction between the moieties (a conformer with the O43−H52⋯O42
and O44−H54⋯O45 IHBs) was chosen as input, to avoid the effect of the removal
of the interaction during the rotation. The scan highlighted five minima (Fig. 2,
considering that angles differing by 360° are the same). Each of the corresponding
geometries was optimized to find the actual geometry corresponding to a given
minimum (as the optimization involves fully relaxed geometry). Since interactions
between S and S′ were not present in the input, this provides information only on
the preferences for the mutual orientations of the moieties and identifies five dif-
ferent combinations of orientations. More conformers were then identified by
changing the IHB patterns in each of the five conformers obtained from the
potential energy profile.
Calculations in solution utilized the Polarizable Continuum Model (PCM, [21–
26]). In this model, the solvent is considered immeasurable and is modelled by a
continuous isotropic dielectric into which the solute is inserted. Thus, the solute
molecule is embedded in a cavity surrounded by the continuum solvent. The
geometry of the cavity follows the geometry of the solute molecule, considering its
solvent accessible surface. The calculations utilised the default settings of Gaus-
sian03 [27] for PCM, namely, Integral Equation Formalism model (IEF, [23–26])
and Gepol model for building the cavity around the solute molecule [28–30], with
simple United Atom Topological Model (UAO) for the atomic radii and 0.200 Å2
for the average area of the tesserae into which the cavity surface is subdivided.
The SCFVAC option was selected to obtain more thermodynamic data.
Calculations in solution were performed as single point (SP) calculations on the
in-vacuo optimised geometries, with the same levels of theory utilised in vacuo. It
was opted to use SP calculations because the size of the molecule makes
re-optimisation in solution computationally expensive. Although SP calculations
cannot provide information on the geometry changes caused by the solvent, they
can provide reasonable information on the energetics, such as the conformers’
relative energies in solution and the energy aspects of the solution process (the free
energy of solvation, ΔGsolv, and its components).
The three solvents considered (chloroform, acetonitrile and water) cover the
ranges of polarity and of hydrogen bonding abilities interesting for biologically
active molecule. Chloroform is an apolar aprotic solvent with low relative per-
mittivity (εr = 4.90) and low dipole moment (µ = 1.04 D [31]). Acetonitrile is a
dipolar aprotic solvent with large relative permittivity (εr = 36.64) and high dipole
moment (µ = 3.92 D [31]). Water is a protic solvent with high relative permittivity
(εr = 78.39) and a sizeable dipole moment (µ = 1.83 D [31]).
Calculations were performed using GAUSSIAN 03, Revision D 01 [27].
All the energy values reported are in kcal/mol and all the distances are in Å.
Acronyms are utilized for the calculation methods and for the media, for con-
ciseness sake on reporting values: HF for HF/6-31G(d,p), DFT for DFT/B3LYP/
6-31+G(d,p), ‘vac’ for vacuum, ‘chlrf’ for chloroform, ‘actn’ for acetonitrile and
‘aq’ for water. Tables with all the numerical values of the properties of the cal-
culated conformers (relative energy, dipole moments, free energy of solvation in the
310 M. K. Bilonda and L. Mammino
solvents considered, HOMO-LUMO energy gaps, etc.), in the results of both cal-
culation methods, and figures showing all the conformers, are included in the
Supplementary Information.
3 Results
Table 1 Letters utilized in the acronyms denoting the conformers of Jozimine A2 in this work,
and their meanings
Letter Meaning Letter Meaning
a Presence of the O43 h Presence of C−H59⋯O46
−H52⋯O42 IHB
b Presence of the O44 t C14−C13−C21−C28 torsion angle
−H54⋯O45 IHB close to +90°
c Presence of the O43−H52⋯π v C14−C13−C21−C28 torsion angle
interaction close to −90°
d Presence of the O44−H54⋯π p C4−C5−C1−C12 torsion angle close
interaction to +90°
e Presence of the O41−H50⋯π q C4−C5−C1−C12 torsion angle close
interaction to −90°
f Presence of the O46−H60⋯π x C22−C23−C31−C38 torsion angle
interaction close to +90°
g Presence of C−H49⋯O41 y C22−C25−C31−C38 torsion angle
close to −90°
Computational Study of Jozimine A2, a Naphthylisoquinoline … 311
The mutual orientations of the four moieties also have significant influence on the
energy of the conformers and, therefore, they are also denoted by specific letters.
Lowercase letters are utilised to provide information about the torsion angles between
moieties, i. e, information about their mutual orientation. The letters ‘p’ and ‘q’ refer to
the torsion angle within the S unit (angle between the A, B and the D, C rings systems,
i.e., C4−C5−C11−C12); the letters ‘x’ and ‘y’ refer to the torsion angle within S′ unit
(angle between the C′, D′ and the A′, B′ rings systems, i.e., C22−C25−C31−C38);
the letters ‘t’ and ‘v’ refer to the torsion angle between the two units S and S′ (i.e., C14
−C13−C21−C28); the meaning of these letters is explained in Table 1. Figure 4
highlights the perpendicularity of the moieties through selected examples. Differently
from michellamine A [12], there is no specific correspondence between the orientation
of the two moieties (within each unit) and the IHB patterns.
When the conformers are mentioned in the text, their acronyms start with JZM,
followed by the letters denoting their characteristics. In tables and figures, ‘JZM’ is not
reported for space-saving reasons. It may be expedient to illustrate how the acronyms
describe the characteristics of individual conformers through some examples.
JZM-a-b-e-f-g-h-p-v-x is a conformer characterized by the presence of the O43
−H52⋯O42 and O44−H54⋯O45 IHBs, the O41−H50⋯π and O46−H60⋯π
interactions and the C−H49⋯O41 and C−H59⋯O46 interactions, with the C4−C5
−C11−C12 torsion angle close to 90°, the C14−C13−C21−C28 torsion angle close
to −90° and C22−C25−C31−C38 torsion angle close to 90°; JZM-c-d-e-f-g-h-q-v-y
is a conformer with the O43−H52⋯π, O44−H54⋯π, O41−H50⋯π and O46
−H60⋯π interactions and the C−H49⋯O41 and C−H59⋯O46 interactions, with the
C4−C5−C11−C12 torsion angle close to −90°, the C14−C13−C21−C28 torsion
angle close to −90° and the C22−C25−C31−C38 torsion angle close to −90°;
JZM-b-c-q-t-x is a conformer characterized by the presence of the O44−H54⋯O45
IHB and the O43−H52⋯π interaction, and with the C4−C5−C11−C12 torsion angle
close to −90°, the C14−C13−C21−C28 torsion angle close to +90° and the C22
−C25−C31−C38 torsion angle close to +90°.
Computational Study of Jozimine A2, a Naphthylisoquinoline … 313
80 conformers were identified, and the frequency calculations confirm that they
correspond to true minima. The relative energies of selected calculated conformers
of JZM in vacuo and in the three solvents considered are reported in Table 2.
Table 2 Relative energies of the selected conformers of jozimine A2 shown in Fig. 3. HF/6-31G
(d,p) and DFT/B3LYP/6-31+G(d,p) results in vacuo and in the solvents considered, respectively
denoted as HF and DFT in the columns’ headings. The results in vacuo are from full optimization
calculations; the results in solution are from single point PCM calculations on the
in-vacuo-optimized geometries. The conformers are listed in order of increasing relative energies
in the HF results in vacuo
Conformer Relative energy (kcal/mol)
HF DFT
vac chlrf actn aq vac chlrf actn aq
a-b-e-f-g-h-p-v-x 0.000 0.000 0.000 0.112 0.000 0.000 0.000 0.095
a-b-e-f-g-h-q-v-x 0.743 0.467 0.278 0.020 0.608 0.404 0.320 0.000
a-b-e-f-g-h-q-t-x 0.793 0.591 0.303 0.243 0.606 0.411 0.290 0.058
a-b-e-f-g-h-q-v-y 1.463 0.637 0.618 0.000 1.202 0.777 0.567 0.005
a-b-e-f-g-h-p-t-x 1.466 0.920 0.707 0.075 1.180 0.825 0.578 0.187
a-b-f-h-q-v-x 5.783 5.411 3.856 1.772 5.634 4.645 3.970 1.860
a-d-e-f-g-h-q-v-x 5.857 5.469 5.429 3.921 5.944 5.799 −a −a
b-c-e-f-g-h-q-t-x 5.934 5.759 5.547 3.902 5.990 5.840 5.679 4.132
a-b-e-g-q-t-x 5.962 4.497 4.003 1.987 5.770 4.757 4.121 1.900
b-c-e-f-g-h-p-v-x 6.209 5.756 5.725 4.095 6.267 6.053 5.902 4.341
a-d-e-f-g-h-p-v-x 6.209 5.756 5.725 4.095 6.267 6.053 5.897 4.375
a-b-e-g-p-v-x 6.269 5.549 4.132 2.037 6.028 4.978 4.266 2.160
a-b-f-h-p-v-x 6.269 4.914 4.192 2.096 6.028 5.004 −a 2.134
a-b-e-g-v-y 6.476 5.168 4.123 1.625 6.202 −a −a 1.819
a-b-f-h-v-y 6.476 5.157 4.037 1.706 6.202 4.917 4.289 1.813
a-b-e-g-p-t-x 6.552 5.316 4.308 1.969 6.204 5.008 4.295 1.999
a-b-f-h-p-t-x 6.552 5.194 4.294 1.945 6.204 5.021 4.278 1.959
c-d-e-f-g-h-p-v-x 11.627 11.590 11.467 8.363 11.508 11.728 11.623 8.798
a-d-e-g-p-t-x 11.682 10.404 9.397 5.723 11.568 10.460 9.692 6.632
a-b-q-v-y 11.899 9.483 7.954 3.038 11.602 −a −a 3.878
a-b-p-t-x 12.096 10.142 8.111 3.817 11.684 9.638 8.295 4.437
c-d-f-h-p-v-x 15.532 14.858 14.216 10.362 15.508 15.181 14.688 10.984
c-d-f-h-q-v-x 15.687 14.974 14.302 3.432 15.648 15.128 14.575 −a
b-c-q-v-y 16.262 13.764 12.531 6.784 16.266 14.080 12.760 7.637
a-d-q-v-y 16.262 13.982 12.425 6.961 16.266 14.099 12.753 7.714
c-d-q-v-y 20.097 18.547 17.013 11.069 20.169 18.675 17.585 11.609
c-d-p-t-x 20.432 18.889 17.427 11.707 20.216 18.933 18.027 12.417
a
The calculation for this conformer did not converge in the given solvent
314 M. K. Bilonda and L. Mammino
The reported conformers comprise the first fourteen lower-energy conformers and
representative conformers of other types (other combinations of IHBs) not
appearing among the first fourteen. The IHB patterns (numbers and types of IHBs
present in a conformer) are the dominant stabilizing factors. Others factors
influencing conformational preferences are the mutual orientation of the two
naphthalene moieties (C, D and C′, D′ ring systems), which determines the mutual
orientation of the two units (S and S′) and the orientation of the two moieties
(naphthalene and isoquinoline) within each unit. The isoquinoline moiety prefers to
be perpendicular to the naphthalene moiety and the two naphthalene moieties
(linked by the inter-units biaryl axis) also prefer to be perpendicular to each another;
this results in the two units (S and S′) preferring to be mutually perpendicular,
which excludes the possibility of an IHB between them.
As mentioned previously, three types of IHBs interactions may be present:
O−H⋯O IHBs (O43−H52⋯O42 and O44−H54⋯O45), O−H⋯π interactions
between the O−H in an isoquinoline moiety and the closest π system in the
naphthalene moiety of the same unit (O41−H50⋯π, O46−H60⋯π) or between an
O−H in the naphthalene moiety of one unit and the closest π system in the
naphthalene moiety of the other unit (O43−H52⋯π, O44−H54⋯π); and C−H⋯O
interactions within an isoquinoline moiety (C−H49⋯O41 and C−H59⋯O46). The
O−H⋯O IHBs have the strongest stabilizing effect. The two O−H⋯O IHBs (O43
−H52⋯O42 and O44−H54⋯O45) have practically the same geometric parameters
in a conformer having both of them simultaneously.
The lowest energy conformers are the conformers having all the IHB-types
interactions simultaneously. The first five lowest energy conformers have relative
energy below 1.5 kcal/mol in both the HF and the DFT results. They have the same
types of IHBs interactions and differ only by the orientations of the moieties.
Altogether, they account for 99.95% of the population in vacuo (Table S12).
The reported X-ray structure of JZM [3] corresponds to the a-b-f-h-p-v-x con-
former, which has 6.269 kcal/mol relative energy and 0.0014 population (Table 2).
This conformer has the two O−H⋯O IHBs (O43−H52⋯O42 and O44
−H54⋯O45), one O−H⋯π interaction and one C−H⋯O interaction. Compared to
the five lowest energy ones (which have all the interactions), this conformer has one
O−H⋯π and C−H⋯O interaction less, which leaves one OH group free. This
phenomenon is observed frequently (e.g., in the case of caespitate [32]): the
crystalline structure differs from the geometry in vacuo by not having one or more
weaker IHBs or other IHB-type interactions that are present in the gas phase.
A possible reason is the need, for molecules in the crystal structure, to have some
donors or acceptors not engaged in intramolecular interactions, so that they are
available for intermolecular interactions with the surrounding molecules.
The symmetry of the molecule results in a number of pairs of conformers with
the same characteristics (IHBs) in the S moiety of one conformer and in the S′
moiety of the other conformer. Pairs of this type will be termed S/S′ symmetric
pairs in the rest of the text.
Computational Study of Jozimine A2, a Naphthylisoquinoline … 315
For the O−H⋯π interactions, a bond length is not defined, as the acceptor is a
whole π system and not an individual atom. It may be convenient to consider the
distance between the H atom of the donor OH and the C atom in the aromatic
system closest to it. The ranges of the distances thus defined are reported in Table 3.
It can be inferred from these ranges that the H⋯C distance is slightly shorter when
the interaction involves an O−H in the naphthalene moiety of one unit and the
closest π system in the naphthalene moiety of the other unit (O43−H52⋯π and O44
−H54⋯π) than when it involves an O−H in the isoquinoline moiety and the closest
π system in the naphthalene moiety of the same unit (O41−H50⋯π and O46
−H60⋯π).
The ranges of the H⋯O distance for the C−H⋯O interactions within the iso-
quinoline moiety (C−H49⋯O41 and C−H59⋯O46) are also reported in Table 3.
This distance is considerably longer than the H⋯O distance for O−H⋯O IHBs,
consistently with the fact that the C−H⋯O interaction is considerably weaker. The
donor–acceptor C⋯O distance for C−H49⋯O41and C−H59⋯O46 is ≈3.0 Å in
both the HF and DFT results. The CĤO bond angle is ≈115.0°/HF and ≈116.0°/
DFT for both C−H49⋯O41 and C−H59⋯O46. Their parameters show that the C
−H⋯O IHBs are weak H-bonds.
Table 4 lists the ranges of the calculated harmonic vibrational frequencies of the
O–H bonds. Only HF frequencies are available in this study because frequency
calculations with the DFT method did not complete (which is probably due to the
high number of atoms in this molecule). When IHBs are present, it is interesting to
consider also the red-shift (lowering of the vibrational frequency of the donor OH)
caused by the IHBs. The red shift is evaluated with respect to the frequency of a
free OH of the same type. Since the OHs in JZM are never free, two model
structures with free OHs and with the other features similar to those of the moieties
in JZM (including a CH3 to mimic the presence of other moieties attached to the
one considered) were used to calculate a reference frequency. These structures are
Computational Study of Jozimine A2, a Naphthylisoquinoline … 317
Table 4 Ranges of the harmonic vibrational frequencies of the OH groups in jozimine A2 and of
the red shifts caused by IHBs
OH Frequency (cm−1) Red shift when engaged in O Red shift when engaged in O−H⋯π
−H⋯O IHB interactions
O41−H50 3735.02−3804.55 − 39.08−64.31
O43−H52 3707.53−3758.99 27.37−60.59 9.13−46.53
O44−H54 3707.73−3759.84 30.03−60.39 8.23−51.28
O46−H60 3736.13−3838.26 − 40.34−65.17
shown in Fig. 5. The ranges of the red shifts of O43−H52 and O44−H54 when
engaged in the different types of IHBs are reported in Table 4. The values of the red
shifts for O−H⋯O IHBs and the O−H⋯π interactions are comparable, which is
consistent with the stabilizing effect of both types of IHBs (although the strength of
the O−H⋯O IHB is expected to be greater than that of the O−H⋯π IHB).
Table 5 reports the values of the ZPE and of the relative energies corrected for
ZPE for the conformers listed in Table 2. The ZPE corrections are very close for all
the conformers. Their values (kcal/mol) are in the 571.572−572.602 kcal/mol
range, with the greater values corresponding to lower energy conformers. The
relative energies corrected for ZPE have the same trends as the uncorrected ones.
Table 6 reports the values of the dipole moment for the conformers listed in
Table 2. The ranges (debye) of the values of the dipole moments in vacuo are 0.50
−77.53/HF and 0.59−77.33/DFT. Both the magnitude and the direction of the
dipole moment vector change according to the conformer and are largely influenced
by the orientation of the OH groups (although the orientation of the aromatic rings
also plays a role).
Conformers with O−H⋯O IHBs have higher dipole moment than conformers
with O−H⋯π interactions, with few exceptions. Conformers with only O−H⋯π
interactions (no O−H⋯O IHBs) are among the conformers with smaller dipole
318 M. K. Bilonda and L. Mammino
Table 5 Relative energy (ΔEcorrect, kcal/mol) corrected for ZPE and ZPE corrections (kcal/mol)
for conformers of jozimine A2 selected among those reported in Table 2. Results from HF/6-31G
(d,p) frequency calculations
Conformer ΔEcorrect ZPE Conformer ΔEcorrect ZPE
correction correction
a-b-e-f-g-h-p-v-x 0.000 572.602 a-b-e-g-p-t-x 6.100 572.150
a-b-e-f-g-h-q-v-x 0.644 572.503 c-d-e-f-g-h-p-v-x 11.036 572.011
a-b-e-f-g-h-q-t-x 0.688 572.497 a-d-e-g-p-t-x 11.053 571.973
a-b-e-f-g-h-q-v-y 1.268 572.407 a-b-q-v-y 11.089 571.791
a-b-e-f-g-h-p-t-x 1.294 572.430 a-b-p-t-x 11.313 571.819
a-d-e-f-g-h-q-v-x 5.545 572.289 c-d-f-h-p-v-x 14.776 571.846
b-c-e-f-g-h-q-t-x 5.616 572.284 c-d-f-h-q-v-x 14.866 571.781
a-b-e-g-q-t-x 5.575 572.215 b-c-q-v-y 15.280 571.621
b-c-e-f-g-h-p-v-x 5.871 572.264 c-d-q-v-y 19.007 571.572
a-d-e-f-g-h-p-v-x 5.872 572.264 c-d-p-t-x 19.433 571.603
moments. The same phenomenon was observed in the previous study of naph-
thylisoquinoline alkaloids with antimalarial activity [16] and in the study of
michellamine A [12]. The mutual orientation of the moieties also has considerable
influence on the dipole moment. Different orientations of the S and S′ units may
cause 1−4 D difference in the dipole moments, and different orientations of the
isoquinoline and naphthalene moieties within each unit may cause 1−2 D difference
in the dipole moments. This trend is reversed with respect to what was observed for
the conformers of michellamine A [12], where the orientation of the units caused a
1−2 D difference and the different orientation of the isoquinoline and naphthalene
moieties within each unit caused a 1−4 D difference. When two conformers of JZM
differ both by the orientation of the S and S′ units and by the orientations of the
isoquinoline and naphthalene moieties, the dipole moment difference is ≈2 D.
Conformers of S/S′ symmetric pairs have the same dipole moment.
Table 7 reports the HOMO-LUMO energy gap for the conformers listed in
Table 2. The gap is influenced by the IHB patterns. Conformers with only one
O−H⋯O IHB and other types of IHBs interactions (O−H⋯π and C−H⋯O), such
as conformers JZM-b-c-e-g-p-v-x, JZM-c-f-h-q-t-x, JZM-a-d-e-g-p-t-x,
JZM-a-d-e-g-q-v-y, have the smallest HOMO-LUMO energy gap. The gap is
slightly greater for conformers with the two O−H⋯O IHBs and other IHB-type
interactions and highest for conformers with only O−H⋯π interactions (which are
accompanied by C−H⋯O interactions if they involve the two moieties within the
same unit). The presence of the C−H⋯O interactions in a conformer slightly
decreases the HOMO-LUMO energy gap with respect to a corresponding con-
former where it is absent; for instance, the gap in JZM-c-d-f-h-p-v-x is ≈1 kcal/mol
less than in JZM-c-d-p-v-x. Conformers of S/S′ have the same HOMO-LUMO
energy.
The estimation of the HOMO-LUMO energy gap shows marked difference
between HF and DFT values. This is a known phenomenon, as DFT substantially
Computational Study of Jozimine A2, a Naphthylisoquinoline … 319
Table 6 Dipole moment of the conformers of jozimine A2 listed in Table 2, in vacuo and in the
three solvents considered. HF/6-31G(d,p) and (DFT/B3LYP/6-31+G(d,p) results, respectively
denoted as HF and DFT in the column headings. The results in vacuo are from full optimization
calculations, the results in solution are from single point PCM calculations on the
in-vacuo-optimized geometries
Conformer Dipole moment (Debye)
HF DFT
vac chlrf actn aq vac chlrf actn aq
a-b-e-f-g-h-p-v-x 2.65 2.67 2.55 2.44 2.81 2.87 2.74 2.58
a-b-e-f-g-h-q-v-x 5.46 5.96 6.06 5.97 5.32 5.93 6.09 5.99
a-b-e-f-g-h-q-t-x 5.03 5.41 5.43 5.34 5.01 5.50 5.57 5.45
a-b-e-f-g-h-q-v-y 6.83 7.46 7.55 7.44 6.59 7.36 7.52 7.39
a-b-e-f-g-h-p-t-x 7.53 8.29 8.50 8.43 7.33 8.28 8.57 8.49
a-b-f-h-q-v-x 4.25 4.95 4.51 4.31 4.04 4.36 4.41 4.22
a-d-e-f-g-h-q-v-x 4.56 4.85 4.87 4.73 4.70 5.11 −a −a
b-c-e-f-g-h-q-t-x 5.03 6.12 6.29 6.28 5.52 6.22 6.46 6.41
a-b-e-g-q-t-x 4.20 4.45 4.47 4.27 4.10 4.39 4.42 4.21
b-c-e-f-g-h-p-v-x 4.11 4.53 4.63 4.69 4.15 4.65 4.81 4.83
a-d-e-f-g-h-p-v-x 4.11 4.53 4.64 4.69 4.15 4.66 4.80 4.83
a-b-e-g-p-v-x 3.93 4.24 4.29 4.37 4.38 4.79 4.85 4.91
a-b-f-h-p-v-x 3.93 4.23 4.28 4.36 4.38 4.78 −a 4.90
a-b-e-g-v-y 4.84 5.08 5.04 4.89 4.36 −a −a 4.27
a-b-f-h-v-y 4.84 5.07 5.04 4.89 4.36 4.59 4.55 4.27
a-b-e-g-p-t-x 4.96 5.21 5.18 4.88 4.78 5.12 5.10 4.78
a-b-f-g-p-t-x 4.95 5.22 5.17 4.88 4.77 5.11 5.09 4.78
c-d-e-f-g-h-p-v-x 0.50 0.20 0.02 0.17 0.70 0.40 0.20 0.21
a-d-e-g-p-t-x 4.62 5.10 5.24 5.20 4.03 4.54 4.70 4.62
a-b-q-v-y 1.38 1.13 −a 0.65 0.59 −a −a 0.41
a-b-p-t-x 1.42 1.19 0.96 0.50 1.08 0.85 0.60 0.15
c-d-f-h-p-v-x 3.83 4.28 4.49 4.65 4.04 4.60 4.82 4.98
c-d-f-h-q-v-x 3.79 4.22 4.39 4.39 3.71 4.56 4.29 −a
b-c-q-v-y 3.91 4.40 4.56 4.67 3.97 4.69 5.00 5.25
a-d-q-v-y 3.91 4.40 4.56 4.67 3.97 4.70 5.00 5.24
c-d-q-v-y 1.28 1.81 2.03 2.40 2.07 2.83 3.19 3.70
c-d-p-t-x 1.33 1.86 2.09 2.64 1.84 2.52 2.87 3.48
a
The calculation for this conformer did not converge in the given solvent
underestimates the values of the gaps [44, 45]. However, the two methods show
similar trends.
Figure 6 shows the shapes of the HOMO and LUMO orbitals for the five lowest
energy conformers, some representative higher energy ones and the conformers of
some S/S′ symmetric pairs. The shapes indicate greater electron density and similar
distribution in the two naphthalene moieties than in the isoquinoline moieties for
Table 7 HOMO-LUMO energy gap of the conformers of jozimine A2 reported in Table 2, in vacuo and in the three solvents considered. HF/6-31G(d,p) and
320
(DFT/B3LYP/6-31+G(d,p) results, respectively denoted as HF and DFT in the column headings. The results in vacuo are from full optimization calculations,
the results in solution are from single point PCM calculations on the in-vacuo-optimized geometries
Conformer HOMO-LUMO energy gap (kcal/mol)
HF DFT
vac chlrf actn aq vac chlrf actn aq
a-b-e-f-g-h-p-v-x 232.762 232.850 232.881 232.919 98.343 98.506 98.563 98.613
a-b-e-f-g-h-q-v-x 253.376 233.929 233.960 234.048 98.908 99.065 99.115 99.203
a-b-e-f-g-h-q-t-x 233.810 233.923 233.967 234.029 98.757 98.946 99.027 99.102
a-b-e-f-g-h-q-v-y 234.921 235.046 235.096 235.159 99.391 99.567 99.642 99.730
a-b-e-f-g-h-p-t-x 234.745 234.883 234.933 234.977 99.579 99.774 99.862 99.931
a-b-f-h-q-v-x 232.341 232.800 232.517 233.007 97.007 97.094 97.389 98.155
a-d-e-f-g-h-q-v-x 224.843 228.733 230.126 231.927 89.031 92.922 −a −a
b-c-e-f-g-h-q-t-x 224.661 228.608 230.158 231.896 88.868 92.752 94.327 96.266
a-b-e-g-q-t-x 232.191 232.272 232.505 233.051 96.963 96.982 97.226 97.923
b-c-e-f-g-h-p-v-x 224.912 228.840 230.359 232.166 89.050 92.991 94.515 96.511
a-d-e-f-g-h-p-v-x 225.157 228.859 230.365 232.134 89.044 92.984 94.515 96.505
a-b-e-g-p-v-x 231.846 231.978 232.235 232.749 96.655 96.718 97.025 97.596
a-b-f-h-p-v-x 231.839 231.971 232.241 232.718 96.655 96.718 −a 97.615
a-b-e-g-v-y 232.938 232.856 232.969 233.709 96.837 −a −a 98.036
a-b-f-h-v-y 232.938 232.850 232.969 233.697 96.843 96.944 97.214 98.036
a-b-e-g-p-t-x 232.498 232.618 232.925 233.640 96.756 96.850 97.163 97.954
a-b-f-h-p-t-x 232.498 232.611 232.900 233.659 96.756 96.850 97.151 97.973
c-d-e-f-g-h-p-v-x 236.558 236.508 236.483 236.433 100.552 100.502 100.489 100.502
a-d-e-g-p-t-x 221.166 225.226 227.064 229.618 85.435 89.257 91.020 93.536
a-b-q-v-y 234.017 234.306 234.506 234.707 98.776 −a −a 99.460
(continued)
M. K. Bilonda and L. Mammino
Table 7 (continued)
Conformer HOMO-LUMO energy gap (kcal/mol)
HF DFT
vac chlrf actn aq vac chlrf actn aq
a-b-p-t-x 234.237 234.462 234.569 234.776 98.889 99.184 99.335 99.567
c-d-f-h-p-v-x 233.051 233.032 233.239 233.747 97.251 97.088 97.295 97.904
c-d-f-h-q-v-x 234.205 233.873 233.967 234.557 97.760 97.835 97.835 0.000
b-c-q-v-y 225.000 229.028 230.591 232.442 88.999 93.016 94.547 96.442
a-d-q-v-y 225.000 229.041 230.559 232.398 89.006 92.978 94.509 96.448
c-d-q-v-y 236.320 236.307 236.289 236.245 100.4077 100.364 100.326 100.282
c-d-p-t-x 236.452 236.401 236.370 236.301 100.3073 100.276 100.257 100.213
a
The calculation for this conformer did not converge in the given solvent
Computational Study of Jozimine A2, a Naphthylisoquinoline …
321
322 M. K. Bilonda and L. Mammino
HOMO
LUMO
Fig. 6 Shapes of the HOMO and LUMO frontier orbitals of the lowest energy conformers of
jozimine A2. HF/6-31G(d,p) results. The initial part of the acronyms is not reported under the
images, because of space reasons; it is the same for all the conformers (JZM)
the five lowest energy conformers—a phenomenon that was also observed for
monomeric naphthylisoquinoline alkaloids with antimalarial activity [16] and for
michellamine A [12]. For higher energy conformers, the electron distribution
becomes considerably different for the two naphthalene moieties. For conformers of
S/S′ symmetric pairs, electrons concentrate on one or the other of the naphthalene
moieties.
Table 2 shows the relative energy of the conformers in the three solvents consid-
ered (chloroform, acetonitrile and water). The relative energy of the conformer
decreases with increasing solvent polarity—which is consistent with common
behaviours. (Acetonitrile has a greater dipole moment than water, but its effect on
solutes is often intermediate between that of chloroform and that of water; therefore,
“increasing solvent polarities” refers to the chloroform-acetonitrile-water sequence
Computational Study of Jozimine A2, a Naphthylisoquinoline … 323
in this discussion). The first five lowest energy conformers have relative energy
below 1 kcal/mol in chloroform and acetonitrile, and below 0.24 kcal/mol in water.
The identification of the lowest energy conformer does not change in chloroform
and acetonitrile with respect to in vacuo, but may be different in water; however, the
relative energies of the five lower-energy conformers are so small in water that a
different identification of the lowest-energy one among them does not affect the
interpretation of the results. Major changes in water occur for some conformers
having relative energy in vacuo greater than 5 kcal/mol. These cases concern
conformers in which O41−H50 or O46−H60 are not engaged in the O−H⋯π
interaction and, therefore, are available to form H-bonds with water molecules.
Although PCM does not take into explicit account solute-solvent H-bonds, it
appears sometimes to take it into account implicitly through the energetics (possibly
as the effect of the point-charges distribution in the areas of the cavity surface
corresponding to an H-bond donor or acceptor) [46]. A confirmation about the
interactions of these conformers with water molecules can be obtained through the
consideration of adducts with explicit water molecules, which might be the object
of a separate study.
The five lowest energy conformers account for 99.89% of the population in
chloroform, 99.66% in acetonitrile and 90.71% in water. The population distribution
of the other conformers (besides the five lowest energy ones) is also different. In
vacuo, no other conformer has a population greater than 0.003%; in chloroform,
there are two conformers with population of 0.02 and 0.01% respectively, and all
the other populations are <0.007%; in acetonitrile, ten other conformers have a
population between 0.02 and 0.05%; in water, three other conformers have popu-
lation between 1.0 and 1.3% and seven conformers between 0.1 and 0.9%.
Conformers of S/S′ symmetric may have slightly different relative energy in
solution, suggesting that the effect of their symmetric situation may somewhat
decrease in solution.
If 3.5 kcal/mol is taken as a cautious threshold value for conformers which
might be responsible for the biological activity, the results in water solution suggest
that most conformers of JZM (all those with relative energy ≤ 3.5 kcal/mol in water
solution) might be considered as potential responsibles for the antimalarial activity
of JZM.
Table 8 reports the solvent effect (free energy of solvation, ΔGsolv) for the
conformers listed in Table 2. ΔGsolv is positive in chloroform and acetonitrile and
negative in water for all the conformers. The values in acetonitrile are considerably
greater than those in chloroform. The electrostatic component of ΔGsolv (Gel) has
negative values in all the three solvents, but considerably more negative in water.
A quick estimation of the octanol/water partition coefficient of JZM (6.3743, [47])
suggests that JZM could be more soluble in non-polar solvents than in water. This
could be due to presence of many aromatic rings and to the high molecular mass.
On the other hand, the negative values of ΔGsolv in water suggest the possibility of
some (although limited) solubility in water or, at least, the possibility that the
molecule may be present also in the water phases of living organisms cannot be
324 M. K. Bilonda and L. Mammino
Table 8 Solvent effect (free energy of solvation, ΔGsolv) of selected conformers of jozimine A2 in
the three solvents considered Results from HF/6-31G(d,p) and DFT/6-31+G(d,p) single point
PCM calculations on the in-vacuo-optimized geometries
Conformer ΔGsolv (kcal/mol)
HF DFT
chlrf actn aq chlrf actn aq
a-b-e-f-g-h-p-v-x 7.19 17.60 −4.12 7.37 18.30 −4.12
a-b-e-f-g-h-q-v-x 6.95 17.66 −4.39 7.29 18.16 −4.64
a-b-e-f-g-h-q-t-x 7.15 17.77 −4.04 7.32 18.14 −4.57
a-b-e-f-g-h-q-v-y 6.81 17.67 −4.70 7.19 17.93 −5.09
a-b-e-f-g-h-p-t-x 7.17 17.89 −4.50 7.22 17.96 −4.87
a-b-f-h-q-v-x 5.76 16.21 −7.62 6.65 16.98 −7.54
a-d-e-f-g-h-q-v-x 6.97 17.86 −5.37 7.61 −a −a
b-c-e-f-g-h-q-t-x 7.19 17.91 −5.46 7.61 18.44 −5.50
a-b-e-g-q-t-x 5.78 16.21 −7.54 6.57 16.93 −7.72
b-c-e-f-g-h-p-v-x 6.84 18.05 −5.26 7.50 18.36 −5.60
a-d-e-f-g-h-p-v-x 7.17 17.95 −5.55 7.50 18.36 −5.57
a-b-e-g-p-v-x 6.38 15.89 −7.96 6.44 16.72 −7.83
a-b-f-h-p-v-x 5.74 15.95 −7.90 6.47 −a −7.86
a-b-e-g-p-t-x 6.01 15.96 −8.09 6.61 16.94 −7.73
a-b-f-h-p-t-x 5.88 15.94 −8.13 6.62 16.93 −7.76
c-d-e-f-g-h-p-v-x 7.57 18.43 −6.35 7.95 18.87 −6.33
a-d-e-g-p-t-x 6.13 16.10 −9.25 6.68 16.93 −8.49
a-b-q-v-y 4.89 −a −12.31 −a −a −11.36
a-b-p-t-x 5.65 14.61 −11.38 5.61 15.32 −10.92
c-d-f-h-p-v-x 6.74 17.09 −8.42 7.30 17.83 −8.25
c-d-f-h-q-v-x 6.80 17.12 −8.53 7.48 18.00 −a
b-c-q-v-y 5.36 15.08 −12.30 5.85 15.53 −11.91
a-d-q-v-y 5.76 15.07 −12.03 5.88 15.54 −11.81
c-d-q-v-y 6.10 15.53 −12.02 6.59 16.50 −11.77
c-d-p-t-x 5.91 15.47 −11.86 6.56 16.70 −11.23
a
The calculation for this conformer did not converge in the given solvent
excluded. It is probable that the ΔGsolv takes into account the presence of OHs,
which favours the solubility in water.
The presence and number of O−H⋯O IHBs, and the O−H⋯π and C−H⋯O
interaction seem to influence ΔGsolv and this effect is different for different solvents.
The first five lowest energy conformers (with a-b-e-f-g-h in common) have the
smallest absolute values of ΔGsolv in water. Conformers in which either O41−H50
or O46−H60 is not engaged in the O−H⋯π IHB have the greatest absolute values
of ΔGsolv in water. Conformers in which only O41−H50 or O46−H60 is engaged
Computational Study of Jozimine A2, a Naphthylisoquinoline … 325
in O−H⋯π interaction have greater values of ΔGsolv than the conformers in which
O41−H50 and O46−H60 are simultaneously engaged in O−H⋯π interaction.
In acetonitrile, the conformers with all the OHs engaged in IHBs have the
greatest values of ΔGsolv. The conformers in which neither O41−H50 nor O46
−H60 are engaged in O−H⋯π IHBs have the smallest absolute values of ΔGsolv
and conformers in which only either O41−H50 or O46−H60 is engaged in the O
−H⋯π IHB have values of ΔGsolv greater than the conformers in which they are not
engaged in them. The behaviour in acetonitrile is different with respect to the
behaviour in water. The importance of solute-solvent H-bonds for acetonitrile may
not be as relevant as in water, both because acetonitrile can only be H-bond
acceptor and because N is a weaker acceptor than O. The trend in chloroform is
almost similar to the trend in acetonitrile, but the conformers having all the possible
O−H⋯π interactions have slightly greater ΔGsolv than the other conformers.
Conformers of S/S′ symmetric pairs may have slightly different ΔGsolv values.
The dipole moments (Table 6) increases in chloroform with respect to in vacuo
and in acetonitrile with respect to chloroform, but decreases slightly in water with
respect to acetonitrile. This might be related to the fact that acetonitrile has greater
dipole moment than water (although several of its effects on solute molecules
appear to be intermediate between those of chloroform and those of water). Con-
formers of S/S′ symmetric pairs have the same value of dipole moment in the
solvents considered.
The HOMO-LUMO energy gap increases with the solvent polarity. Conformers
with all the O−H⋯π and C−H⋯O interactions (c-d-e-f-g-h) have the highest values
of the HOMO-LUMO energy gap in all the media, and conformers with only one O
−H⋯O IHB (O43−H52⋯O42 or O44−H54⋯O45) have the lowest energy
gap. The range of the HOMO-LUMO energy gap in vacuo is mostly 221.17
−236.56/HF and 85.43−100.79/DFT. The gap mostly increases in solution, but it
decreases in some cases; the change with respect to in vacuo is −0.41−4.34/HF and
−0.25−4.15/DFT in chloroform; −0.43−6.04/HF and −3.39−1.84/DFT in ace-
tonitrile and −0.15−8.56/HF and −0.13−8.34/DFT in water. Conformers of S/S′
symmetric pairs may have slightly different HOMO-LUMO gaps.
moieties. The O−H⋯O IHBs are the dominant factor determining conformational
preferences and energetics.
The low relative energy of many conformers in water solution suggests that a
comparatively high number of conformers may be potential responsibles for the
molecule’s biological activity. The values of the free energy of solvation suggest
greater solubility in water than in the other two solvents considered in this study,
which would be consistent with the high polarity of the JZM molecule, but not with
the octanol/water partition coefficient of JZM or with its high molecular mass.
It is interesting to note that (differently from a number of other molecules), JZM
appears to respond to the three solvents considered according to their increasing
dipole moment, so that the responses to water are slightly smaller than those to
acetonitrile.
A typical feature of this molecule is the presence of axial chirality (stereoiso-
merism resulting from the non-planar arrangement of four groups in pairs about a
chirality axis). The presence of axial chirality suggests that rotation about the C13
−C21 bond is not free because of steric hindrance. A space-filling visualisation with
Chem3D [47] appears to confirm the hindrance, at least for lower energy con-
formers. On the other hand, the rotation scans about this bond (and about the other
two biaryl bonds) were not disrupted (i.e., the steric hindrance did not appear in the
bond-rotation model). A possible reason is the fact that the rotation scan input was a
conformer with 12.811 kcal/mol relative energy (2.32 × 10−8 population);
although expedient to identify minima in the orientation of the moieties, this
conformer is poorly populated and, therefore, it does not contribute to the mole-
cule’s actual (or experimentally determined) behaviour.
An analysis of the differences with the behaviours and properties of michel-
lamine A may be useful for a better understanding of the difference in the biological
activities of these largely similar molecules.
References
1. World Health Organization (2014) World malaria report 2014 summary. http://www.who.int/
malaria/publications/world_malaria_report_2014/wmr-2014-no-profiles.pdf
2. World Health Organization (2016) World malaria report 2016. http://www.who.int/malaria/
media/world-malaria-report-2016/en/
3. Bringmann G, Zhang G, Büttner T, Bauckmann G, Kupfer T, Braunschweig H, Brun R,
Mudogo V (2013) Chem Eur J 19:916–923
4. Bierer DE, Dener JM, Dubenko LG, Gerber RE, Litvak J, Peterli S, Peterli-Roth P,
Truong TV, Mao G, Bauer BE (1995) J Med Chem 38:2628
5. Ganellin CR, Mitchell RC, Young RC (1988) In: Melchiorre C, Giannella M (eds) Recent
advances in receptor chemistry. Elsevier Science Publishers B.V., Amsterdam, pp 289–306
6. Guha S, Majumdar D, Bhattacharjee AK (1992) J Mol Struct (Theochem) 256:61
7. Sjoberg P, Murray JS, Brinck T, Evans P, Politzer P (1990) J Mol Graph 8:81
Computational Study of Jozimine A2, a Naphthylisoquinoline … 327
1 Introduction
the issue regarding the Gâteaux differentiability of density functionals [6, 7]. How-
ever, the Fréchet differentiability of density functionals remained unresolved until
Lindgren and Salomonson claimed its plausibility recently [8–10]. In this paper, we
reexamine Lindgren and Salomonson’s analysis to gain a better understanding about
the Fréchet differentiability of density functionals in DFT.
Mathematically speaking, a functional is a mapping from a function to a number.
G[f ], a functional of function f (x), can be expressed as f (x) ↦ G[f ]. The differential
of a functional, dG[f , 𝛿f ], is the part of the difference,
where 𝛿G∕𝛿f (x) is the functional derivative of G[f ] with respect to f at point x. For
the sake of brevity, ⟨⋅⟩ is adopted as a shorthand notation for integration throughout
the text.
In DFT, there are two kinds of functional derivative: the Gâteaux derivative and
the Fréchet derivative [6–10]. Following Lindgren and Salomonson [8–10], all the
density functionals are defined on a convex space of densities,
√
= {𝜌 | 𝜌 ≥ 0, 𝜌 ∈ 1 (3 )} , (3)
exists, it is called the Gâteaux differential of G at 𝜌0 in the direction 𝛿𝜌. If the limit
exists for any 𝛿𝜌 such that 𝜌0 + 𝛽𝛿𝜌 ∈ , we say G is Gâteaux differentiable at 𝜌0 .
Stated in another way, provided the functional G is Gâteaux differentiable at 𝜌0 ,
the functional difference upon a density variation, 𝜌0 (𝐫) → 𝜌0 (𝐫) + 𝛽𝛿𝜌(𝐫), has two
terms,
𝛿G[𝜌0 , 𝛽𝛿𝜌] = G[𝜌0 + 𝛽𝛿𝜌] − G[𝜌0 ] = 𝛽dG[𝜌0 , 𝛿𝜌] + R[𝜌0 , 𝛽𝛿𝜌] , (6)
Functional Derivatives and Differentiability in Density-Functional Theory 333
R[𝜌0 , 𝛽𝛿𝜌]
lim =0, (7)
𝛽→0+ 𝛽
in which 𝛿G∕𝛿𝜌0 (𝐫) is called the Gâteaux derivative of G at 𝜌0 (𝐫). Similar to a direc-
tional derivative, the Gâteaux derivative is a functional of 𝜌0 (𝐫) only, although along
various directions it might take different expressions (e.g., different functions of the
variable 𝐫).
If the last term of the functional difference,
⟨ ⟩
𝛿G
𝛿G[𝜌0 , 𝛿𝜌] = G[𝜌0 + 𝛿𝜌] − G[𝜌0 ] = 𝛿𝜌(𝐫) + R[𝜌0 , 𝛿𝜌] , (9)
𝛿𝜌0 (𝐫)
instead satisfies
R[𝜌0 , 𝛿𝜌]
lim =0, (10)
||𝛿𝜌||→0 ||𝛿𝜌||
for the norm of 𝛿𝜌(𝐫), ||𝛿𝜌|| = ⟨|𝛿𝜌(𝐫)|⟩, 𝛿G∕𝛿𝜌0 (𝐫) is called the Fréchet derivative.
The Fréchet derivative is a global derivative: all directions approaching 𝜌0 (𝐫) yield
the same derivative (with the same expression). By default, Fréchet differentiability
is stronger than Gâteaux differentiability.
∑
N
∑
N
∑N
̂ + V̂ = − 1
Ĥ = T̂ + 𝜔W ∇2i + 𝜔
1
+ v(𝐫i ) , (11)
2 i=1 i<j
|𝐫i − 𝐫j | i=1
334 P. Xiang and Y. A. Wang
where i and j are dummy electron indices. For convenience, the first two operators in
Eq. (11) can be grouped into a single Hohenberg-Kohn (HK) universal operator [2],
F̂ 𝜔 = T̂ + 𝜔W
̂ . (12)
In Eqs. (11) and (12), when 𝜔 = 0, we have the noninteracting system; when 𝜔 = 1,
we have the fully interacting system instead. It is then straightforward to show that
the Schrödinger equation governing this system,
̂ = Ev Ψ ,
HΨ (13)
where the subscript “N − 1” indicates the integration to be carried out over all spatial
and spin coordinates except for one spatial coordinate of a single electron.
In Eq. (16), the normalization constraint of the electron density can be relaxed to
allow ⟨Ψ|Ψ⟩ ≠ 1. As a result, the potential-energy density functional V[𝜌] and the
HK universal density functional F 𝜔 [𝜌] are defined in the domain of unnormalized
densities:
̂
V[𝜌] = ⟨Ψ|V|Ψ⟩ = ⟨v(𝐫) 𝜌(𝐫)⟩ , (17)
and
F 𝜔 [𝜌] = FLL
𝜔
[𝜌] = inf ⟨Ψ|T̂ + 𝜔W|Ψ⟩
̂ , (18)
Ψ→𝜌
where “inf ” is the infimum or the greatest lower bound and the subscript “LL”
denotes the Levy-Lieb functional [5, 18]. It should be noted that the density of con-
cern, 𝜌, belongs to , not the convex set of N-representable densities, N :
Functional Derivatives and Differentiability in Density-Functional Theory 335
√
N ≡ {𝜌 | 𝜌 ≥ 0, ⟨𝜌⟩ = N, 𝜌 ∈ 1 (3 )} . (19)
𝛿Ts [𝜌]
= 𝜀i − veff (𝐫), (21)
𝛿𝜌(𝐫)
̂
⟨Ψ|F̂ 𝜔 |Ψ⟩ = ⟨Ψ|H|Ψ⟩ − ⟨v(𝐫) 𝜌(𝐫)⟩
= E0 ⟨Ψ|Ψ⟩ + ⟨𝛿Ψ|Ĥ − E0 |𝛿Ψ⟩ − ⟨v(𝐫) 𝜌(𝐫)⟩
⟨[ ] ⟩
E0
= − v(𝐫) 𝜌(𝐫) + ⟨𝛿Ψ|Ĥ − E0 |𝛿Ψ⟩ , (23)
N
𝜔
FLL [𝜌0 + 𝛿𝜌] = inf ⟨Ψ|F̂ 𝜔 |Ψ⟩
Ψ→𝜌0 +𝛿𝜌
⟨[ ] ⟩
E0 [ ]
= − v(𝐫) 𝜌0 (𝐫) + 𝛿𝜌(𝐫) + inf ⟨𝛿Ψ|Ĥ − E0 |𝛿Ψ⟩ . (24)
N Ψ0 +𝛿Ψ→𝜌0 +𝛿𝜌
𝜔 𝜔 𝜔
𝛿FLL = FLL [𝜌0 + 𝛿𝜌] − FLL [𝜌0 ]
⟨[ ] ⟩
E0
= − v(𝐫) 𝛿𝜌(𝐫) + inf ⟨𝛿Ψ|Ĥ − E0 |𝛿Ψ⟩ . (25)
N Ψ0 +𝛿Ψ→𝜌0 +𝛿𝜌
Then, the condition for Fréchet differentiability requires the last term in Eq. (25) to
satisfy
⟨𝛿Ψ|Ĥ − E0 |𝛿Ψ⟩
inf → 0 , as ||𝛿𝜌|| → 0 , (26)
Ψ0 +𝛿Ψ→𝜌0 +𝛿𝜌 ||𝛿𝜌||
for all density variations in the neighborhood of 𝜌0 (𝐫). Lindgren and Salomonson
argued that Eq. (26) was plausible because the numerator is quadratic in 𝛿Ψ whereas
𝛿𝜌 is only linear in 𝛿Ψ [8], and they analyzed this issue further based on their proof
of the Gâteaux differentiability of the Levy-Lieb functional [9].
When the ground state is degenerate, it can no longer be represented by a PS-
v-representable density. Instead, we should use an ensemble v-representable (E-v-
representable) density [6, 21],
∑ ∑
𝜌0 = sk 𝜌k0 , sk ≥ 0 , sk = 1, (27)
k k
where 𝜌k0 is a PS-v-representable density of the Hamiltonian in Eq. (11). For concise-
ness, we use the same notation 𝜌0 to represent either a PS-v-representable density or
an E-v-representable density [6, 21], and whenever needed, we will specify the type
of the density. For this degenerate case, by using a similar method as above, Lindgren
and Salomonson [9] has also shown the plausibleness of the Fréchet differentiability
of the Lieb functional [5],
∑ ⟨ ⟩
| ̂ || Ψk ,
FL𝜔 [𝜌] = inf sk Ψk |T̂ + 𝜔W (28)
sk ,Ψ →𝜌
k | |
k
where {Ψk } is any set of orthonormal eigenfunctions of the Hamiltonian in Eq. (11).
where N is the total number of variables and m is the number of variables excluded
in the integration. If f is chosen to be the seed function, we can define the order-m
SOF space that consists all functions of order-m strongly orthogonal to the same seed
function f .
Because spin operator commutes with the Hamiltonian, they share a common
complete set of eigenfunctions. In the forthcoming discussion, we only consider
wavefunctions in Hilbert space spanned by this complete set. Let the seed function
be the N-electron GS wavefunction Ψ0 and the integration in Eq. (29) be over all spa-
tial and spin coordinates except m spatial coordinates, and label as its associated
order-1 SOF space and as its order-0 SOF space with excluded. For a general
Hamiltonian (0 ≤ 𝜔 ≤ 1), there is no doubt that exists. Since we integrate over
all spin coordinates, if any non-GS eigenfunction Ψ has a different spin multiplicity
from that of Ψ0 , we have
⟨Ψ0 |Ψ⟩N−1 = 0 , (30)
the Fréchet derivative does not exist. For the wavefunction Ψp , its corresponding
electron density is
⟨ ⟩
|
𝜌p (𝐫) = N Ψp |Ψp
|
⟨ ⟩ N−1 ⟨ ⟩ (⟨ ⟩ )
= N Ψ0 |Ψ0 N−1 + 𝛽 2 N Ψ ||Ψ N−1 + 2𝛽NRe Ψ ||Ψ0 N−1
|
= 𝜌0 (𝐫) + 𝛽 2 𝜌 (𝐫) , (32)
( )
where Re ⟨Ψ |Ψ0 ⟩N−1 , the real part of ⟨Ψ |Ψ0 ⟩N−1 , is zero because of the nature
of Ψ . When 𝛽 approaches 0, 𝜌p (𝐫) also approaches 𝜌0 (𝐫). Clearly, 𝜌p (𝐫) lies in the
neighborhood of 𝜌0 (𝐫) within . Equation (32) specifies the density variation path
for the forthcoming discussion. For later convenience, we label as the set of all
legitimate 𝜌p (𝐫) for a given Ψ0 or 𝜌0 (𝐫). Throughout the text, Re(⋅) will be used to
denote the real part of the quantity involved.
̃ be a trial wavefunction yielding the
Along this particular variational path, let Ψ
same density 𝜌p (𝐫):
Ψ̃ = Ψ0 + 𝜆Ψt ⟼ 𝜌p (𝐫) , (33)
338 P. Xiang and Y. A. Wang
̃ Ψ⟩
𝜌̃(𝐫) = N⟨Ψ| ̃ N−1
( )
= N⟨Ψ0 |Ψ0 ⟩N−1 + 𝜆2 N⟨Ψt |Ψt ⟩N−1 + 2𝜆NRe ⟨Ψ0 |Ψt ⟩N−1
( )
= 𝜌0 (𝐫) + 𝜆2 𝜌t (𝐫) + 2𝜆NRe ⟨Ψ0 |Ψt ⟩N−1 . (35)
Substituting Eqs. (32) and (35) into Eq. (37), one has
(⟨ ⟩)
𝜆2 ⟨𝜌t (𝐫)⟩ + 2𝜆NRe Ψ0 ||Ψt = 𝛽 2 ⟨𝜌 (𝐫)⟩ . (38)
Since ⟨𝜌 (𝐫)⟩ and ⟨𝜌t (𝐫)⟩ are all equal to N, Eq. (38) can be readily simplified to
∞
(⟨ ⟩) ∑ ( ⟨ ⟩)
2 2 | 2
𝛽 = 𝜆 + 2𝜆Re Ψ0 |Ψt = 𝜆 + 2𝜆 Re ci Ψ0 ||Ψi = 𝜆2 + 2𝜆c0 . (39)
i=0
At any specific point along the variational path, the value of 𝛽 is fixed, and we can
solve 𝜆 in terms of 𝛽 based on Eq. (39):
√
𝜆 = −c0 ± c20 + 𝛽 2 . (40)
Because both 𝜆 and 𝛽 approach 0 concurrently near the end of the variational path,
only the positive sign (see Appendix 3) is allowed in Eq. (41). Thus, as 𝛽 → 0, we
get
1 2 1
𝜆= 𝛽 − 3 𝛽4 + ⋯ . (42)
2c0 8c0
Immediately, we can conclude that towards the end of the variational path, 𝜆 is of
the same magnitude as 𝛽 2 ∕c0 .
If c0 = 0, Eq. (39) immediately reduces to 𝜆2 = 𝛽 2 . In this case, the wavefunction
variation 𝛿Ψ in Eq. (26) can be regarded as linear in 𝛽 and the density variation 𝛿𝜌 =
𝛽 2 𝜌 is quadratic in 𝛽 as shown by Eq. (32). This immediately invalidates Fréchet
differentiability by destroying Eq. (26). Nonetheless, we are able to gain much deeper
understanding about the structure of Ψt through the following analysis.
Consider the case when c0 ≠ 0. Because of Eqs. (32), (35), and (36), we obtain
(⟨ ⟩ )
𝜆2 𝜌t + 2𝜆NRe Ψ0 ||Ψt N−1 = 𝛽 2 𝜌 . (43)
Substituting Eq. (39) into Eq. (43) and grouping terms according to the powers of 𝜆,
we have
[ 0
]
∑ (⟨ ⟩ )
2𝜆 c0 𝜌 − N |
ci Re Ψ0 |Ψi N−1 = 𝜆2 (𝜌t − 𝜌 ) , (44)
i
where the summation on the left-hand side (LHS) is only within the combined space
of and Ψ0 ,
0 ≡ ∪ Ψ0 . (45)
Separating the Ψ0 contribution from the summation, we can simplify Eq. (44) to
∑ (⟨ ⟩ ) ( )
2c0 (𝜌0 − 𝜌 ) + 2N ci Re Ψ0 ||Ψi N−1 = 𝜆 𝜌 − 𝜌t . (46)
i
Because 𝜌0 , 𝜌 , and all Ψi in Eq. (46) have no dependence on 𝜆, we must admit that
ci corresponding to any wavefunction in 0 must be a function of 𝜆.
Based on perturbation theory, ci (𝜆) can be expanded as
ci (𝜆) = c(0)
i
+ c(1)
i
𝜆 + h.o. , (47)
where “h.o.” represent higher-order terms in 𝜆. It can be readily shown that all the
c(0)
i
must be zero in order to satisfy Eq. (46) for any 𝜆 → 0, whereas c(1)
i
can be zero
or non-zero. Consequently, all {ci } for Ψi in 0 vary at least linearly in 𝜆. Some ci
340 P. Xiang and Y. A. Wang
̃ H|
⟨Ψ| ̃ = ⟨Ψ0 + 𝜆Ψt |H|Ψ
̂ Ψ⟩ ̂ 0 + 𝜆Ψt ⟩
̂ 0 ⟩ + ⟨Ψ0 + 𝜆Ψt |H|𝜆Ψ
= ⟨Ψ0 + 𝜆Ψt |H|Ψ ̂ t⟩
̃ 0 ⟩ + ⟨Ψ0 |H|𝜆Ψ
= E0 ⟨Ψ|Ψ ̂ 2 ̂
t ⟩ + 𝜆 ⟨Ψt |H|Ψt ⟩
̃ Ψ⟩
= E0 ⟨Ψ| ̃ − E0 ⟨Ψ|𝜆Ψ
̃ 2 ̂
t ⟩ + E0 ⟨Ψ0 |𝜆Ψt ⟩ + 𝜆 ⟨Ψt |H|Ψt ⟩
̃ Ψ⟩
= E0 ⟨Ψ| ̃ − 𝜆2 E0 ⟨Ψt |Ψt ⟩ + 𝜆2 ⟨Ψt |H|Ψ
̂ t⟩
( )
2 2 ̂
= (1 + 𝛽 )E0 + 𝜆 ⟨Ψt |H|Ψt ⟩ − E0 . (48)
( )
̂ t ⟩ − E0 in Eq. (48)
To achieve this task, one only needs to minimize 𝜆2 ⟨Ψt |H|Ψ
under the following two constraints:
∞
∑
c2i = 1 , (49)
i=0
and
𝜌̃(𝐫) = 𝜌p (𝐫) . (50)
The second constraint, Eq. (50), is equivalent to the following equation based on our
previous analysis:
[
]
c0 (𝜌0 − 𝜌 ) ∑ (⟨ ⟩ )
2𝜆 + ci Re Ψ0 ||Ψi N−1
N i
( ∞
)
𝜌 ∑ ⟨ ⟩
=𝜆 2
− ci cj Ψi ||Ψj N−1 . (51)
N i,j
and
( )
̂ t ⟩ − E0 − h𝐀 − ⟨g(𝐫) 𝐁⟩
𝛀 = 𝜆2 ⟨Ψt |H|Ψ
(∞ )
∑
= 𝜆2 ̂ j ⟩ − E0 − h𝐀 − ⟨g(𝐫) 𝐁⟩
ci cj ⟨Ψi |H|Ψ
i,j
( ∞
)
∑
2
=𝜆 ci cj Ej 𝛿ij − E0 − h𝐀 − ⟨g(𝐫) 𝐁⟩
i,j
[ ∞
]
∑
2
=𝜆 c2i (Ei − E0 ) − h𝐀 − ⟨g(𝐫) 𝐁⟩ , (54)
i=1
where h and g(𝐫) are the Lagrange multipliers corresponding to the two constraints
in Eqs. (52) and (53), and Eq. (49) has been used to derive the last expression. Min-
imizing Eq. (54) with respect to {ci }, one obtains
⟨[ ] ⟩
𝜌0 − 𝜌 ∑∞ (⟨ ⟩ )
𝜆 +𝜆 cj Re Ψ0 ||Ψj N−1 g(𝐫) = −hc0 , (55)
N j=0
and
⟨[ ] ⟩
(⟨ ⟩ ) ∞
∑ (⟨ ⟩ )
𝜆 Re Ψi || Ψ0 N−1 + 𝜆 cj Re Ψi ||Ψj N−1 g(𝐫)
j=0
[ 2
]
= 𝜆 (Ei − E0 ) − h ci (for i ≠ 0) . (56)
Because c0 is at least linear in 𝜆 as we showed above, we can readily infer from Eq.
(55) that g(𝐫) must take the following form,
∞
∑ g(k) (𝐫)
g𝜆 (𝐫) = g(0) (𝐫) + 𝜆k , (57)
k=1
k!
where “h.o.” denotes higher-order terms in 𝜆. Obviously, we reach the same conclu-
sion as we derived before: {ci } for Ψi ∈ is at least linear in 𝜆 towards the end of the
variational path. For those {ci } for Ψi ∈ , where Ψi is order-1 strongly orthogonal
to Ψ0 , the first term in the square brackets on the LHS of Eq. (56) disappears, and
Eq. (56) reduces to
⟨ ⟩ ⟨ ⟩
∑0 (⟨ ⟩ ) ∑ (⟨ ⟩ )
𝜆2 cj Re Ψi ||Ψj N−1 g(𝐫) + 𝜆2 cj Re Ψi ||Ψj N−1 g(𝐫)
j j
2
= 𝜆 (Ei − E0 )ci − hci . (59)
For this equation to be valid at 𝜆 → 0, the LHS and the right-hand side (RHS) should
have the same dependence on 𝜆. On the RHS, the first term decays faster than the
second term, and the second term will dominate when 𝜆 approaches 0. Therefore,
we must match the magnitude of the second term on the RHS to the LHS. Of course,
we cannot match it with the second term on the LHS because doing so will lead to
self inconsistency. Then, the second term on the RHS must decay the same way as
the first term on the LHS. So, {ci } for Ψi ∈ are proportional to 𝜆3 or higher-than-
cubic terms in 𝜆. Unfortunately, such a behavior is contradictory to the normalization
∑
constraint in Eq. (49). Otherwise, i c2i will become 0 as 𝜆 → 0. Hence, we conclude
∑
that this contradiction must come from the initial assumption: Ψt = i ci Ψi , where
the expansion is over the complete set of eigenfunctions of H.̂
To resolve this contradiction, we have to modify our assumption about Ψt . We
∑
notice that if the summation i ci Ψi includes any wavefunction from 0 , the same
problem will persist. Thus, Ψt can only be expanded in ,
∑
Ψt = ci Ψi . (60)
i
Integrating both sides of Eq. (61) over the entire space of 𝐫, one finds
𝜆2 = 𝛽 2 , (62)
where all traces of 𝜆 (or 𝛽) are completely gone. The minimization will yield the
optimal set of expansion coefficients {̄ci }, which has no dependence on 𝜆 or 𝛽 from
the appearance of Eq. (64). At this stage, we are ready to test the condition for Fréchet
differentiability shown in Eq. (26):
where Eq. (62) is used to simplify the expression after the second equal sign. Evi-
dently, Eq. (65) suggests that the condition for Fréchet differentiability is not fulfilled.
Since the above infimum approaches a nonzero constant towards the end of the
variational path, it is possible to combine the linear-order term of 𝛿𝜌 from the sec-
ond term with the first term in Eq. (25), and the new resulting residual term might
satisfy the condition for Fréchet differentiability. However, the corresponding path-
dependent functional derivative will be different from the one obtained by Lindgren
and Salomonson [8–10]. This is contradictory to the fact that the Fréchet derivative
is a global derivative, independent of variational path. Therefore, we have no choice
𝜔
but to conclude that the Levy-Lieb density functionals FLL are not Fréchet differ-
entiable at PS-v-representable densities. The places where Fréchet differentiability
breaks down (e.g., along our specially designed variational path in ) are exactly the
same locations where the Gâteaux derivative espouses different forms.
In the above analysis, we worked within unnormalized density domain. The situa-
tion in normalized density domain is almost identical. This can be straightforwardly
proven by normalizing the wavefunction after finishing the minimization process
in the unnormalized wavefunction space, simply because all the wavefunctions of
our concern are normalizable in Hilbert space and the normalization factor does not
affect the expectation values of observables. In Appendix 2, we have offered a much
more detailed but rather lengthy proof to further confirm our assessment here.
After showing the non-Fréchet differentiability of the Levy-Lieb functionals, we
are ready to examine the differentiability of the Lieb functionals, shown in Eq. (28).
344 P. Xiang and Y. A. Wang
Using similar derivations as shown in Sect. 2, we get the variation of the Lieb func-
tional:
⟨[ ] ⟩
∑ E0 ∑
𝛿FL [𝜌0 ] = sk − v(𝐫) 𝛿𝜌k (𝐫) + sk k inf ⟨𝛿Ψk |Ĥ − E0 |𝛿Ψk ⟩
k
N k
Ψ0 +𝛿Ψk →𝜌k0 +𝛿𝜌k
⟨[ ] ⟩
E0 ∑
= − v(𝐫) 𝛿𝜌0 (𝐫) + sk k inf ⟨𝛿Ψk |Ĥ − E0 |𝛿Ψk ⟩ , (66)
N k
Ψ0 +𝛿Ψk →𝜌k0 +𝛿𝜌k
based on Eq. (27), where the total density variation 𝛿𝜌 is a linear combination of
individual variation of 𝜌k0 ,
∑
𝛿𝜌 = sk 𝛿𝜌k , (67)
k
where sk is the same as that in Eq. (27). Then, the condition for the Lieb functional
to be Fréchet differentiable is
̃ k = Ψ k + 𝜆k Ψ k .
Ψ (71)
0 t
By the same analysis, we can show that each scaling parameter 𝜆k must satisfy
𝜆2k = 𝛽 2 , (72)
and each trial wavefunction Ψkt can only be expanded in the corresponding space,
∑
Ψkt = cki Ψki . (73)
i
Functional Derivatives and Differentiability in Density-Functional Theory 345
Consequently, we have
which is positive and cannot be zero for any positive semi-definite set of {sk } because
Therefore, the condition for Fréchet differentiability of the Lieb functional is not
fulfilled, and the Lieb functional is not Fréchet differentiable at any E-v-representable
densities.
where the scaling parameter 𝛽 can take any real value as long as the resultant density
𝜌(𝐫) is not negative,
𝜌(𝐫) = 𝜌0 (𝐫) + 𝛽𝜂(𝐫) ≥ 0 , (77)
everywhere in the entire space, for a given 𝜂(𝐫). As said before, if we let 𝜂(𝐫) to be
a density in , then 𝛽 must be in the range [0, 1] because of the convexity of .
Perdew and Levy proposed a kind of unconventional density variations (UDVs)
which satisfy [23]:
𝜂(𝐫)
√ ∉ 2 . (78)
𝜌0 (𝐫)
This class of UDVs gives rise to an unconventional energy variation, 𝛿Ev , of sub-
quadratic order of 𝛽 [25],
𝛿Ev [𝜌0 , 𝛿𝜌] = Ev [𝜌0 + 𝛿𝜌] − Ev [𝜌0 ] ∝ O(𝛽 b ), 1 < b < 2 , (79)
where “h.o.” encompasses all higher-order terms in 𝛿𝜌(𝐫). Zhang and Wang [25]
recently discovered that Eq. (80) cannot be expanded to second order in 𝛿𝜌(𝐫) for
the UDVs of the first kind [23]. In other words, the second (or any higher-order)
functional derivative of the total energy functional may not exist for all allowed den-
sity variations within the current DFT framework. This imposes a serious problem
in DFT since the effort to interpret the second functional derivative as the chemical
hardness [1] will fail for the UDVs of the first kind [23]. Some modification of the
density variation domain must be in place to rescue the situation.
From a different perspective, we introduce a new definition of UDV based on
our discussion in Sect. 3. If a density variation, 𝛿𝜌(𝐫) = 𝜌(𝐫) − 𝜌0 (𝐫), comes from a
density 𝜌(𝐫) ∈ , then the density variation is an UDV. For these UDVs of the second
kind, their corresponding Gâteaux derivatives are different from the conventional
one [8–10], or simply, not Fréchet derivatives. √
Let us write the wavefunction in Eq. (31) differently as Ψ′p = Ψ0 + 𝛽Ψ , the
associated density variation will become linear in 𝛽,
From Eq. (82), we know that the second term on the RHS of Eq. (83) is of linear
order in 𝛿𝜌′(𝐫) when the density is approaching 𝜌0 (𝐫). We can then split the second
term in Eq. (83) into a first-order term of 𝛿𝜌′(𝐫) and a higher-order residual:
R[𝜌0 , 𝛿𝜌′ ]
lim =0. (85)
𝛽→0 𝛽
If a CDV path is chosen instead, we will get the Gâteaux derivative previously
obtained by Lindgren and Salomonson [8–10]:
𝜔 |
𝛿FLL | E0
| = − v(𝐫) . (88)
𝛿𝜌(𝐫) ||𝜌 N
0
Equations (87) and (88) clearly indicate that the nonuniqueness of the Gâteaux
derivative along different paths. For the Lieb functionals FL𝜔 , it can be easily shown
that the same conclusion is still valid.
Because of the existence of UDVs of the first kind [23] and the second kind,
density variation domain has to be cleansed so that consistent results can be obtained.
However, there remains one problem: What is the relationship between these two
kinds of UDVs? In other words, does the set of Perdew and Levy’s UDVs contains
our UDVs or is the opposite true? Consider a noninteracting two-electron atom with
nuclear charge Z = 2, its ground state is 1s2 and the GS wavefunction is denoted by
Φ1s2 . If both electrons are excited to the 2s orbital, we would get a wavefunction Φ2s2
from the space of Φ1s2 . The density variation,
with 0 ≤ 𝛽 ≤ 1, would belong to the UDVs of the second kind. Due to the asymptotic
behavior of wavefunctions [1], we have
√
lim 𝜌1s2 (𝐫) ∼ e−2𝐫 2I1s2
(90)
𝐫→∞
and √
lim 𝜌2s2 (𝐫) ∼ e−2𝐫 2I2s2
, (91)
𝐫→∞
348 P. Xiang and Y. A. Wang
where I1s2 and I2s2 are first ionization potentials for 1s2 and 2s2 configurations,
respectively. Because the system concerned here is a noninteracting one, it can be
readily shown that I1s2 = 4I2s2 . In this case, we have
𝜂(𝐫) 𝜌 2 (𝐫) √ √
lim √ = lim √ 2s = lim e−𝐫 ( 8I2s2 − 2I1s2 ) = 1 . (92)
𝐫→∞
𝜌0 (𝐫) 𝐫→∞ 𝜌1s2 (𝐫) 𝐫→∞
Consequently, Perdew and Levy’s condition for the UDVs of the first kind, Eq. (78),
is satisfied and the density variation, Eq. (89), also belongs to the UDVs of the first
kind.
However, not all UDVs of the second kind are UDVs of the first kind. Here is
an example. Replacing the atom in the last example with a noninteracting 4-electron
hydrogen-like atom and letting 𝛿𝜌 to be from the wavefunction Φ1s2 3s2 , which is in
the space of GS wavefunction Φ1s2 2s2 , then we have a new density variational path:
𝜌′ (𝐫) = 𝜌0 (𝐫) + 𝛽𝛿𝜌(𝐫) = 𝜌1s2 2s2 (𝐫) + 𝛽𝜌1s2 3s2 (𝐫) , (93)
𝜂(𝐫) 𝜌 2 2 (𝐫) √ √
lim √ = lim √ 1s 3s = lim e−𝐫 ( 8I1s2 3s2 − 2I1s2 2s2 ) = 0 , (94)
𝐫→∞
𝜌0 (𝐫) 𝐫→∞ 𝜌1s2 2s2 (𝐫) 𝐫→∞
because the first ionization potentials of the two configurations, 1s2 2s2 and 1s2 3s2 ,
satisfy a different equation:
Clearly, the condition for Perdew and Levy’s UDVs cannot be fulfilled in this case.
From the discussion above, we can see that the set of the UDVs of the second
kind is not enclosed in the set of the UDVs of the first kind. However, the question
of whether the UDVs of the second kind fully contain the UDVs of the first kind is
still left open. Most likely, these two sets of UDVs share some common elements,
but not mutually inclusive.
The current definition of density variation domain [25] is based on the pioneer
work of Lieb [5] and of Englisch and Englisch [6, 7, 24]. For wavefunctions in
Hilbert space, the density of concern belongs to the convex set of N-representable
densities, N . For wavefunctions in Fock space, the density domain is the direct sum
of N : ⨁
≡ N , (96)
N∈+
without the UDVs of the second kind, where N is defined in Appendix 2. In other
words, the density must be in the nexus of the above two restricted density domains:
{ | }
| 𝜌 − 𝜌
𝜌(𝐫) ∈ N′
≡ N ∩ ≡ 𝜌 || 𝜌0 ∈ N , 𝜌 ∈ N , √
′N 0 2
∈ , 𝜌 ∉ N ,
| 𝜌0
|
(99)
for wavefunctions in Hilbert space. Accordingly, we have to restrict the density for
wavefunctions in Fock space:
⨁
𝜌(𝐫) ∈ ′ ≡ N′ . (100)
N∈+
5 Conclusions
Within the current framework of DFT, the Levy-Lieb functionals are not Fréchet
differentiable at PS-v-representable densities and the Lieb functionals are not Fréchet
differentiable at E-v-representable densities. For the Levy-Lieb functionals, when
the density variation comes from a wavefunction in , the Gâteaux derivatives will
become path-dependent, taking a different form from the conventional ones. For the
Lieb functionals, when the variation of each individual PS-v-representable density
that comprises the total density comes from a wavefunction in its corresponding
space, the Gâteaux derivatives will take a different form from the conventional ones.
Based on the analysis on UDVs, we have proposed necessary modifications on the
density variational domain on which density functionals are Fréchet differentiable
and possess the conventional analytic density expansion through second order in
density variation.
Acknowledgements Financial support for this project was provided by a grant from the Natural
Sciences and Engineering Research Council (NSERC) of Canada.
350 P. Xiang and Y. A. Wang
Appendix 1
Here, we will briefly introduce some mathematical concepts relevant to our discus-
sion in the main text. All the following content are adopted from an introductory
book on functional analysis [26].
Definition 1 A vector space is a set of elements called vectors with two opera-
tions called addition and scalar multiplication, which satisfy the following axioms.
Definition 2 If x and y are two points of a vector space, then the line segment joining
them is the set of elements {𝛽x + (1 − 𝛽)y | 0 ≤ 𝛽 ≤ 1}. A subset S of a vector space
is convex if the line segment of joining any two points in S is contained in S.
Definition 3 Let and be two vector spaces with the same system of scalars.
Then a function (or mapping) that maps uniquely the elements of onto elements
of ,
T∶ → (101)
whenever
dX (x, x0 ) < 𝛿 . (103)
Definition 9 A complete (with respect to the norm) normed vector space is called
a Banach space.
The smallest value of K which satisfies this inequality is denoted by ||T|| and called
the norm of T. It can be verified that this norm for operators satisfies the axioms for a
norm function and that we may therefore talk of the vector space of bounded linear
transformations T ∶ → . This normed vector space is denoted by (, ).
but is in general neither linear nor continuous in s. Nor does the existence of the
Gâteaux differential at x ensure continuity of T at x. For example,
{
𝜉13
(𝜉1 , 𝜉2 ≠ 0)
f (𝜉1 , 𝜉2 ) = 𝜉2 . (107)
0 (𝜉1 = 𝜉2 = 0)
At point (0, 0), it can be easily shown that the Gâteaux differential exists and it is
zero. Clearly, the Gâteaux differential is a continuous linear operator. However, f is
not continuous at (0,0). Therefore, we cannot relate the Gâteaux differentiability of
T to the continuity of T.
Let us go forward on the basis that is also a normed vector space. Suppose
dT(x; s) is linear and continuous in s for some x ∈ , then we may write
where 𝜖∕𝜆 → 0 as 𝜆 → 0 with x and s fixed. However, the convergence may not be
uniform with respect to s and in that case T cannot be approximated by a linear oper-
ator with uniform accuracy in the neighborhood of x. If we further demand uniform
convergence then we arrive at the strong derivative.
with
||𝜖(x; s)||
lim =0. (112)
||s|| →0 ||s||
The operator TF′ (x) is called the Fréchet or strong derivative of T at x. The Fréchet
derivative at x is unique. It can be shown that the existence of the Fréchet derivative
of T at x implies continuity of T at x.
Theorem 1 If the Gâteaux derivative TG′ (x) exists in the neighborhood of x and is
continuous with respect to the norm in (, ) at x, then the Fréchet derivative
TF′ (x) exists and is equal to TG′ (x).
Appendix 2
In this appendix, we show that the Fréchet derivative does not exist in the normalized
density domain, N .
Define a normalized path wavefunction,
√
Ψp = 1 − 𝛽 2 Ψ0 + 𝛽Ψ , (113)
√ √ ∞
∑
̃=
Ψ 1 − 𝛽 2 Ψ0 + 𝜆Ψt = 1 − 𝛽 2 Ψ0 + 𝜆 ci Ψi ⟼ 𝜌p (𝐫) , (115)
i=0
following form:
̃ Ψ⟩
𝜌̃(𝐫) = N⟨Ψ| ̃ N−1
√ ( )
= (1 − 𝛽 2 )N⟨Ψ0 |Ψ0 ⟩N−1 + 𝜆2 N⟨Ψt |Ψt ⟩N−1 + 2𝜆 1 − 𝛽 2 NRe ⟨Ψ0 |Ψt ⟩N−1
√ ( )
= (1 − 𝛽 2 )𝜌0 (𝐫) + 𝜆2 𝜌t (𝐫) + 2𝜆 1 − 𝛽 2 NRe ⟨Ψ0 |Ψt ⟩N−1 . (116)
At any point, the trial density is identical to the path density to ensure that the density
variation is actually along the path we designed:
Therefore, we have ⟨ ⟩
𝜌(𝐫)⟩ = 𝜌p (𝐫) .
⟨̃ (118)
Substituting Eqs. (114) and (116) into Eq. (118) and simplifying the result, one
derives √
𝛽 2 = 𝜆2 + 2𝜆c0 1 − 𝛽 2 . (119)
At one specific point on the variational path, the value of 𝛽 is fixed, we can solve 𝜆
in terms of 𝛽 based on Eq. (119):
√ √
𝜆 = −c0 1 − 𝛽 2 ± c20 (1 − 𝛽 2 ) + 𝛽 2 . (120)
√ [ √ ]
1 2
𝜆 → −c0 1 − 𝛽 2 ± c0 1 − 𝛽 2 + 𝛽 +⋯ . (121)
2c0
Again (see Appendix 3), the positive sign is chosen in Eq. (121), and we have
1 2
𝜆→ 𝛽 + ⋯ , as 𝛽 → 0 . (122)
2c0
Immediately, we can conclude that towards the end of variational path, 𝜆 is of the
same magnitude of 𝛽 2 ∕c0 . In other words, 𝜆 also approaches zero at nearly the same
rate as 𝛽 2 ∕c0 approaches zero.
Because of Eqs. (114), (116), and (117), we obtain
√ (⟨ ⟩ )
𝜆2 𝜌t + 2N𝜆 1 − 𝛽 2 Re Ψ0 ||Ψt N−1 = 𝛽 2 𝜌 . (123)
Functional Derivatives and Differentiability in Density-Functional Theory 355
where the summation on the LHS is only within 0 . At 𝛽 → 0, we find that the
coefficients {ci } for Ψi ∈ 0 are linear in 𝜆.
After knowing the property of {ci } for wavefunctions in 0 , we then investigate
other remaining {ci } for wavefunctions in . At one particular point on the varia-
tional path (𝛽 fixed), we optimize trial wavefunction to find out the set of coefficients
{ci } that yields the lowest energy for
⟨√ ⟩
̃ H| ̃ =
̂ Ψ⟩ | |√
⟨Ψ| 1 − 𝛽 2 Ψ0 + 𝜆Ψt |Ĥ | 1 − 𝛽 2 Ψ0 + 𝜆Ψt
| |
[ √ ]
= E0 − 𝛽 − 2𝜆c0 1 − 𝛽 2 E0 + 𝜆2 ⟨Ψt |H|Ψ
2 ̂ t⟩
( )
̂ t ⟩ = E0 + 𝜆2 ⟨Ψt |H|Ψ
= E0 − 𝜆2 E0 + 𝜆2 ⟨Ψt |H|Ψ ̂ t ⟩ − E0 , (125)
where Eq. (119) has been used to simplify the expression after the second equal sign.
Obviously, we only need to minimize the last term in Eq. (125) under the following
two constraints:
∞
∑
c2i = 1 , (126)
i=0
and
𝜌̃(𝐫) = 𝜌p (𝐫) . (127)
The density constraint, Eq. (127), is equivalent to the following identity based on
our previous analysis:
[
]
√ c0 (𝜌0 − 𝜌 ) ∑ (⟨ ⟩ )
2𝜆 1 − 𝛽2 + |
ci Re Ψ0 |Ψi N−1
N i
( )
𝜌 ∑∞ ⟨ ⟩
|
= 𝜆2 − cj ci Ψj |Ψi . (128)
N | N−1
i,j
∞
∑
𝐀= c2i − 1 , (129)
i=0
356 P. Xiang and Y. A. Wang
[
]
√ c (𝜌 − 𝜌 ) ∑ (⟨ ⟩ )
𝐁 = 2𝜆 1 − 𝛽 2 0 0
+ ci Re Ψ0 ||Ψi N−1
N i
( ∞
)
2 𝜌 ∑ ⟨
|
⟩
−𝜆 − ci cj Ψi |Ψj N−1 , (130)
N i,j
and
( )
̂ t ⟩ − E0 − h𝐀 − ⟨g(𝐫) 𝐁⟩
𝛀 = 𝜆2 ⟨Ψt |H|Ψ
(∞ )
∑
= 𝜆2 ̂ j ⟩ − E0 − h𝐀 − ⟨g(𝐫) 𝐁⟩
ci cj ⟨Ψi |H|Ψ
i,j
( ∞
)
∑
2
=𝜆 ci cj Ej 𝛿ij − E0 − h𝐀 − ⟨g(𝐫) 𝐁⟩
i,j
[ ∞
]
∑
2
=𝜆 c2i (Ei − E0 ) − h𝐀 − ⟨g(𝐫) 𝐁⟩ , (131)
i=1
where h and g(𝐫) are the Lagrange multipliers corresponding to the two constraints
in Eqs. (129) and (130). Minimizing Eq. (131) with respect to {ci }, one obtains
⟨[ √ ] ⟩
1 − 𝛽 2 (𝜌0 − 𝜌 ) ∑∞ (⟨ ⟩ )
𝜆 +𝜆 cj Re Ψ0 ||Ψj N−1 g(𝐫) = −hc0 , (132)
N j=0
and
⟨[ ] ⟩
√ (⟨ ⟩ ) ∞
∑ (⟨ ⟩ )
𝜆 1 − 𝛽 2 Re Ψi || Ψ0 N−1 + 𝜆 cj Re Ψi ||Ψj N−1 g(𝐫)
j=0
[ ]
= 𝜆2 (Ei − E0 ) − h ci (for i ≠ 0) . (133)
Substituting Eq. (134) into Eq. (133) and ignoring the higher-order terms as 𝜆 → 0,
we obtain an equation for Ψi ∈ ,
√ ⟨ (⟨ ⟩ ) ⟩
−hci = 𝜆 1 − 𝛽 2 Re Ψi || Ψ0 N−1 g(0) (𝐫) + h.o. , (135)
Functional Derivatives and Differentiability in Density-Functional Theory 357
where “h.o.” denotes higher-order terms in 𝜆. Therefore, we reach the same conclu-
sion as before: {ci } for Ψi ∈ is linear in 𝜆 towards the end of variational path.
For those {ci } for Ψi ∈ , utilizing the additional fact that Ψi is order-1 strongly
orthogonal to Ψ0 , we can further simplify Eq. (133) to
⟨ ⟩ ⟨ ⟩
∑0 (⟨ ⟩ ) ∑ (⟨ ⟩ )
𝜆2 cj Re Ψi ||Ψj N−1 g(𝐫) + 𝜆2 cj Re Ψi ||Ψj N−1 g(𝐫)
j j
2
= 𝜆 (Ei − E0 )ci − hci . (136)
For this equation to be valid at 𝜆 → 0, the LHS and the RHS must have the same
dependence on 𝜆. On the RHS, the first term decays faster than the second term, and
the second term will dominate when 𝜆 approaches 0. Therefore, we must match the
magnitude of the second term on the RHS to the LHS. Of course, we cannot match
it with the second term on the LHS because doing so will lead to self inconsistency.
Then, the second term on the RHS must decay in the same way as the first term
on the LHS. Thus, {ci } for Ψi ∈ are proportional to 𝜆3 . Unfortunately, such a
𝜆3 -behavior is contradictory to the normalization constraint in Eq. (126), because
∑ 2
i ci will become 0 as 𝜆 → ∑ 0. Hence, we conclude that this contradiction must come
from the assumption: Ψt = i ci Ψi , where the expansion is over the complete set of
eigenfunctions of H.̂
To resolve the contradiction, we have to modify our assumption about the expan-
∑
sion of Ψt . We notice that if the summation i ci Ψi includes any wavefunction from
0 , the same problem will persist. Therefore, Ψt can only be expanded in ,
∑
Ψt = ci Ψi . (137)
i
Integrating both sides of Eq. (138) over the entire space, one obtains
𝜆2 = 𝛽 2 , (139)
Suppose this minimization will yield the optimal set of expansion coefficients, {̄ci },
which have no dependence on 𝜆 and 𝛽 from the appearance of Eq. (141). Then, we
have
where Eq. (139) is used to simplify the expression after the third equal sign. Evi-
dently, Eq. (142) suggests that the condition for Fréchet differentiability proposed
by Lindgren and Salomonson [8–10] is not fulfilled. In other words, the Fréchet
derivative does not exist in the normalized density domain N either.
Appendix 3
In this appendix, we analyze the consequence of choosing the negative sign in Eqs.
(41) and (121). In the end, we will conclude that this particular choice is fully equiv-
alent to the more natural decision made in the main text and Appendix 2.
Let us start from a unified version of Eqs. (40) and (120):
√
𝜆 = −ac0 ± a2 c20 + 𝛽 2 , (143)
√
where constant a = 1 and 1 − 𝛽 2 in the main text and Appendix 2, respectively.
Obviously, if c0 = 0 or c0 → 0 as 𝛽 → 0, both 𝜆 and 𝛽 approach 0 concurrently near
the end of the variational path.
Functional Derivatives and Differentiability in Density-Functional Theory 359
𝜆 = −2ac0 − 𝜆′ , (144)
which immediately suggests that as 𝛽 → 0, (𝜌0 − 𝜌t ) and the coefficients {ci } for
Ψi ∈ are linear in 𝜆′ . Because 𝜌t → 𝜌0 , Ψt → c0 Ψ0 with |c0 | → 1, as 𝛽 → 0.
Therefore, at the end of the variational path (𝛽 = 0 and 𝜆′ = 0), 𝜆 = −2ac0 , Ψt =
̃ = −aΨ0 .
c0 Ψ0 , |c0 | = 1, and Ψ
Evidently, the choice of the negative sign in Eqs. (41) and (121) yields a fully
equivalent, alternative trial wavefunction,
̃ ′ = −aΨ0 − 𝜆′ Ψt ,
Ψ (147)
References
1. Parr RG, Yang W (1989) Density-functional theory of atoms and molecules. Oxford University
Press, New York
2. Hohenberg P, Kohn W (1964) Phys Rev 136:B864
3. Kohn W, Sham LJ (1965) Phys Rev 140:A1133
4. Wang YA, Xiang P (2013) In: Wesolowski TA, Wang YA (eds) Recent advances in orbital-free
density functional theory, Chap. 1. World Scientific, Singapore, pp 3–12
5. Lieb EH (1983) Int J Quantum Chem 24:243
6. Englisch H, Englisch R (1983) Phys Stat Sol 123:711
7. Englisch H, Englisch R (1984) Phys Stat Sol 124:373
8. Lindgren I, Salomonson S (2003) Phys Rev A 67:056501
9. Lindgren I, Salomonson S (2003) Adv Quantum Chem 43:95
10. Lindgren I, Salomonson S (2004) Phys Rev A 70:032509
11. Ekeland I, Temam R (1976) Convex analysis and variational problems. North-Holland, Ams-
terdam
12. Harris J, Jones RO (1974) J Phys F 4:1170
360 P. Xiang and Y. A. Wang
Jean Maruani
Abstract In previous papers, we revisited the Dirac equation and conjectured that
the electron can be viewed as a massless charge spinning at light speed, this internal
motion being responsible for the rest mass involved in external motions and
interactions. Implications of this concept on basic properties such as time, space,
electric charge, and magnetic moment were considered. The present paper inves-
tigates the deviations of the resulting gyromagnetic factor, fine-structure constant,
and gravitational invariant from their integer approximates, and their implication in
a better understanding of the electromagnetic, gravitational, and other interactions.
In previous papers [1–4], we revisited the Dirac equation and developed a model for
the electron, making conjectures about its rest mass, spin motion, effective size, and
electric charge. This led us to investigate universal constants related to this model.
In the present paper, after recalling specificities related to the Dirac equation, we
elaborate on the light shed by peculiarities of these constants on the relations
between the electromagnetic and gravitational interactions.
J. Maruani (✉)
Laboratoire de Chimie Physique-Matière et Rayonnement,
CNRS & UPMC, 4 Place Jussieu, #32-42, 75005 Paris, France
e-mail: [email protected]
Today, the Dirac equation can be derived from more general theoretical frame-
works [5–8]. But the inductive derivation originally given by Dirac [9, 10] has shown
great heuristic value: it has explained the spin kinetic momentum and magnetic
moment, predicted antimatter, and set the ground for quantum electrodynamics.
In the 1920s, light appeared alternatively as geometric rays following Fermat’s
principle of least optical path (stemming from Snell-Descartes’ laws for reflection
and refraction), or as electromagnetic waves obeying Maxwell’s differential equa-
tions, or as massless particles following Planck-Einstein’s quantum relations:
E = hν = hc ̸ λ, p = E ̸ c; ð1Þ
E = mc2 . ð2Þ
λ B = h ̸ p, p = mv. ð3Þ
T = p2 ̸2m, V = ke e2 ̸ r , ð5Þ
and were not Lorentz-invariant. This hybrid character was partly corrected by
adding ad hoc spin-symmetry conditions for the resulting eigenfunctions Ψ .
The Dirac Electron and Elementary Interactions … 363
A further step was taken by Klein and Gordon, who used as Hamiltonian H the
full energy: mc2 = (m20 c2 + p2)1/2 c, where p2 = p21 + p22 + p23 with pi = mvi along
xi, m = m0γ, γ = ð1 − v2 ̸c2 Þ1 ̸2 . This led to an equation that was relativistic (it had
time and space operators on the same footing), but not quantic (it was quadratic in
the time operator and did not allow to apply the probability principle). Dirac’s feat
was to design a road towards an equation that was both symmetric and linear in
time and space operators, by introducing 4-D matrices multiplying the momentum
operators.
Dimensionwise, an invariant ‘momentum’ p0 ≡ m0c can be defined for a par-
ticle at rest, and an overall ‘momentum’ p4 ≡ mc related to the time coordinate x4
≡ ct. With these notations, it can be written:
p4 2 = p0 2 + p1 2 + p2 2 + p3 2 . ð6Þ
This expression for the invariant (rest mass) ‘momentum’ p0 is similar to that for
the invariant (proper interval) ‘coordinate’ x0:
x4 2 = x0 2 + x1 2 + x2 2 + x3 2 . ð7Þ
Using the notations recalled in Eq. (8) (note that there is no coordinate derivative
associated with the invariant momentum p0), the Schrödinger (9), Klein-Gordon
(10), and Dirac (11) equations can be written as:
It can be seen that the Schrödinger equation, being linear in p4 but quadratic in
the pi’s, is rather a diffusion equation, while the Klein-Gordon equation reduces to a
wave equation for p0 = 0. In the Dirac equation, the αμ matrices are independent of
the p’s and x’s as well as Hermitian and normalized. For Eq. (11) to be equivalent
to Eq. (10), these matrices must also be four-dimensional and anticommutative.
There results that any vector representative of an eigenfunction Ψ must have
four components or, alternatively, that Ψ contains a variable that may take on four
values. Dirac explained that these are the well-known two components of the spin
(±½) and in addition positive and negative values for the mass energy (±mc2). The
existence of a spin kinetic momentum thus appears unseparable from negative-
energy states: both stem from the matrix linearization of a wave equation involving
a quadratic form for the energy.
364 J. Maruani
Before the Dirac equation elucidated its origin, the electron spin had entered
quantum mechanics in two different ways [13]. (1.) A non-energetic, symmetry
requirement for systems of identical particles (Pauli 1925), antisymmetry of the
wave function, this implying an internal dynamical variable with two possible
values. (2.) The deflection of the trajectories of silver atoms by an inhomogeneous
magnetic field (Stern and Gerlach 1922) and the splitting of the spectral lines of
atoms by a magnetic field (Goodsmit and Uhlenbeck 1925), which implied an
intrinsic magnetic moment interacting with the field. The electron spin magnetic
moment is responsible for most of the macroscopic magnetism, from the oxygen
that we breathe to the hard disks of our computers.
To have the spin magnetic moment show up, Dirac made it interact with a
magnetic field. And to have its spin kinetic momentum appear, he had it combined
with an orbital kinetic momentum [10]. Equation (10) was thus extended to include
interactions with an electromagnetic potential ðA4 , AÞ:
h i
ðp4 + e A4 ̸ cÞ − α0 p0 − α ⋅ ðp + e AÞ Ψ = 0. ð12Þ
Note that the invariant momentum p0 is not affected by the external potential ðA4 , AÞ.
Writing H = m0c2 + H′, Dirac showed that, to first order:
In addition to the classical potential and kinetic energies, there appears an extra term,
which he interpreted as due to the interaction of the magnetic field B with an intrinsic
magnetic moment: μs = − ðeℏ ̸ 2m0 Þσ = − μB σ, μB being the Bohr magneton.
The spin kinetic momentum does not give rise to any potential energy. To show
its existence, Dirac computed the angular momentum integrals for an electron
moving in a central electric field (e.g., that of a nucleus):
H = p4 c = − e A4 ðrÞ + c α0 p0 + c α ⋅ p. ð14Þ
Dirac interpreted this as meaning that the electron has a spin kinetic momentum:
s = ðℏ ̸ 2Þ σ, which is to be added to the orbital kinetic momentum l to get a constant
of the motion. The directions of s and μs being defined by the same matrix vector σ,
one has:
μs = − ðe ̸ m0 Þs. ð16Þ
The spin magnetic moment μs differs from the orbital magnetic moment μl by a
factor 2:
as if the ‘loop’ described by the electron in its spin motion had half the length of its
orbital ‘loop’. The total magnetic moment can then be written:
where we introduce the dimensionless factor ge, whose deviations from 2 (the value
given in Dirac’s theory) will be discussed in this paper. This factor distinguishes the
spin magnetic moment from the orbital magnetic moment derived classically. Being
the ratio of μs (in units of μB) to s (in units of ℏ), it is called gyromagnetic factor.
In another computation, Dirac used a field-free Hamiltonian to determine at
which velocity the electron ‘spins’ to acquire kinetic and magnetic momenta [10]:
H = cðα0 p0 + α1 p1 + α2 p2 + α3 p3 Þ. ð19Þ
Making use of the properties of the αk’s he obtained, for any component vk of the
electron velocity:
iℏ ∂2 αk ̸ ∂t 2 = 2 ð∂ αk ̸ ∂tÞ H. ð21Þ
This differential equation can be integrated twice, yielding the explicit time
dependence of the velocity and then of the position. One first obtains:
νe = 2 mc2 ̸ h. ð23Þ
The constant part gives the average velocity through a time interval larger than
ν−1
e , which is observed in practical measurements, while the oscillatory part explains
why the instantaneous velocity has eigenvalues ±c. Further integration yields the
time dependence of the electron coordinate xk, and it appears that the Zitterbewegung
amplitude is of the order of the Compton radius: rC =−λC ̸ 2 = λC ̸ 4π .
Thus, in addition to its external motion (e.g., an orbital motion around a
nucleus), governed by de Broglie’s wavelength λB, the electron is endowed with an
internal motion (Zitterbewegung), governed by Compton’s wavelength λC. Schrö-
dinger showed that Zitterbewegung vanishes when one takes expectation values
over wave packets made up entirely of positive or negative energy states. This was
understood by de Broglie [15] as it resulting from a wave beat between the two
coupled matter and antimatter energy states, the beat frequency νe being the dif-
ference of the two frequencies. This oscillation may be pictured as a Lissajous curve
Lπ ̸2, 1: 2 .
Then, the average mass of the vibrating entity can be considered as null,
departures from this value being allowed by Heisenberg’s uncertainty principle, i.e.:
The value ½ of the spin of the electron (Eq. 15) and the related value 2 of its
gyromagnetic factor (Eq. 13) stem from the beat frequency νe of the positive and
negative energy waves being twice the electron mass-energy frequency (Eq. 23).
However, in Dirac’s theory, the electron interacts with an electromagnetic field
that fulfills relativistic but not quantum requirements. Further consistency was
reached by quantizing the electromagnetic field, which led to quantum electrody-
namics (QED) [7]. Resulting zero-point field (ZPF) oscillations entail ‘radiative
corrections’ that are responsible for the Lamb shift between the 2s1/2 and 2p3/2
levels of hydrogenoid atoms [13] and for the departure of ge from the Dirac integer
value 2. Several authors have used Feynman diagrams to compute increasingly
accurate corrections to ge, yielding the following expansion [18]:
ge ðmeasuredÞ ≈ 2.002 319 304 362; ge ðcorrectedÞ ≈ 2.000 000 000 110 ð60Þ.
This was interpreted by endowing the Dirac ‘point charge’ with a tiny but finite
size: ρe ∼ 10−22 m, much smaller than the electron classical radius: r0 ≈ 2.82 ×
10−15 m, but larger than the Planck length: lP ≈ 1.62 × 10−35 m. It would be that
tiny charge which undergoes Zitterbewegung in a range set by the Compton radius:
rC ≈ 1.93 × 10−13 m.
This discussion deals solely with the free electron interacting with an applied
field. For electrons bound in paramagnetic systems, the measured values of ge are
effective values including, to second order, the orbital momentum and spin-orbit
coupling [20]. The effective factor measured may then be a tensor ge, whose
principal values and axes will depend on the anisotropy of the molecule or of the
crystal site bearing the unpaired electron. Then the extreme accuracy of the value
measured for ge on the free electron is a sign of the extreme isotropy of vacuum
fluctuations.
368 J. Maruani
α ≡ ke e2 ̸ℏ c ≡ 1 ̸ a ≈ 1 ̸ 137.0359991. ð26Þ
The dimensionless quantity a was identified with the ratio of the electron
velocity in the first Bohr orbit to that of light. Its inverse a was given a more general
significance by Eddington [22], who proposed the integer value 137 on speculative
grounds. In atomic units, a measures the velocity of light c; and in Planck units, the
electron charge e = qP a2. Then a can be seen as expressing the strength of the
electromagnetic interaction between charged particles.
The fine-structure constant α (or its inverse, the electric parameter a) shows up in
various domains in physics. For instance, it occurs (together with π) in the
expansion of εg given in Eq. (24). In earlier papers [1–4], we recalled that the
electron Compton diameter (or reduced wavelength) 2rC (−λC ) is the geometric
average of the classical electrostatic radius: r0 = kee2/m0c2, and the hydrogen Bohr
radius: a0 = ℏ2 ̸ ke me e2 , the ratio of this harmonic relation being α:
It is commonly believed that Z = 137 sets a limit for the periodic table. This is
due to the fact that, when using the point-nucleus model, the energy expression
includes a factor: [1 − (aZ)2]1/2, which becomes imaginary for Z > 137. However,
this model is a poor approximation for superheavy elements. If one estimates the
1s orbital radius in these elements and compares it with actual nuclear sizes, it
displays a strong overlap with the nucleus, and the factor above appears as an
artifact [23]. There are several models of a finite nuclear charge distribution, none
of which gives this factor [24]. Therefore, there may in principle be elements with
Z > 137, although they might be too unstable to be observed. Nevertheless, the fact
that Z = 137 is a limit for point nuclei remains a peculiarity of a.
Another reservation about the number 137 is that a is close to this value only
when it is measured in our low-energy world. At a W boson energy (≈ 81 Gev),
a decreases to about 128; and at grand-unification energy, it merges with similar
constants defining the strong and weak nuclear forces: ae ≈ aw ≈ as [25]. But, here
again, the fact that a ≈ 137 at the limit of low energy remains a peculiarity of a
[26].
The Dirac Electron and Elementary Interactions … 369
For the gyromagnetic factor ge, both the integer value 2 resulting from the Dirac
equation and the shift of measured values from 2 have theoretical explanations: the
integer value is ultimately the expression of the duality matter/antimatter, and the
corrections listed in Eq. (24) express the interactions with the quantum vacuum. For
the electric parameter a, no such understanding is available, neither for the integer
value 137 nor for the measured shift of ∼ 0.26 ppt (Eq. 26). Feynman referred to
the constant a in these terms:
It has been a mystery ever since it was discovered more than fifty years ago […]. You
would like to know where this number for a coupling comes from: is it related to π or
perhaps to the base of natural logarithms? Nobody knows; it is one of the greatest damn
mysteries of physics: a magic number that comes to us with no understanding by man. You
might say the ‘hand of God’ wrote that number […]. We know what kind of a dance to do
experimentally to measure this number very accurately; but we don’t know what kind of
dance to do on the computer to make this number come out without putting it in secretly!
Richard Feynman, The Strange Theory of Light and Matter (Princeton UP, 1985).
Over the years, a kind of mystic has developed about the peculiar value a ≈ 137,
as illustrated in the quotation below:
The fine-structure constant holds a special place among cult numbers: [it] seduces otherwise
sedate people into seeking mystical truths and developing uncollaborated theories …. The
cult of 137 began with scientists who already had quite a reputation, including Pauli, Jung,
Heisenberg, and most notably Eddington […]. In more recent years, the tradition has spread
to the larger community of science theory hobbyists ….
Robert Munato, http://www.mrob.com/pub/num/n-b137_035.html.
However, although Eddington’s speculations about the integer 137 are dis-
putable, this number has, even more than 2, a number of mathematical properties
that make it rather unlikely that its physical occurrences be merely due to chance. In
addition to being the 33rd prime number, it is a superadditive prime in base 10
(1 + 3 + 7 = 11 and 1 + 1 = 2). It is also a Chen prime (twinned with 139), a
Stern prime (the 4th among 8), an Euler prime, an Eisenstein prime, a Pythagorean
prime, a binary quadratic prime, and an optimal primeval prime (the 10th of the 15
primes generated with its 3 digits).
It has been shown [27] that a and 137 are related to the conspicuous numbers π
and e (related themselves through the magic relation: eiπ = −1), and to the har-
monic sum 137/60 and golden ratio Φ (defined from the equation: Φ = Φ − 1 + 1),
through the relations (precise within ∼ 0.24 ppm, 0.15 ppm, 0.33 ppt and 0.15 ppt,
respectively):
̸ 60
a2 ≈ 1372 + π 2 ; Log a ̸ Log 137 ≈ a2 ̸ a2 − 1 ; 3 a ≈ 60 Φ4 ; Φ137 ≈ 3.
Sanchez [27] has also disclosed puzzling occurrences of a (or 137) in the mass
distributions of nucleotide bases and amino-acids in terms of the hydrogen mass
mH. For instance, every amino-acid being coded by three couples of nucleotide
bases (in pairs A-T and G-C), the average mass M C of all codons appears to be
(within <0.2%):
370 J. Maruani
M C ̸ mH ≈ 33 a! ð28Þ
The number 137 was known to the Egyptians, and 60 to the Chaldeans [4]. Both
occur in the 5th sum of the harmonic series: Σ n ð1 ̸ nÞ, i.e.: Σ 5 = 137/60. 137 is also
a central polygonal number within the series Cn = n(n + 1)/2 + 1. More precisely,
one has: C2 = 4, C4 = 11, C16 = 137, C60 = 1831, then 42 + 112 = 137 and [27]:
2
1372 + 18312 ≈ 36 π 10 ≈ 1836.122 ≈ mp ̸ me ð1836.152 Þ. ð29Þ
The number 137 also relates to the Catalan series built on the Mersenne numbers:
Mn = 2n − 1. The sequence of its terms is: M2 = 3, M3 = 7, M7 = 127, and (the
ending term) M127 ≈ 1.701411835 × 1038 [28]. Now M’7 ≡ M2 + M3 + M7 =
137 ≈ a (Eq. 26), and M’127 yields (within <0.6%) the Hubble radius RU of the
Universe in terms of the Compton radius rC of the electron [27]:
′
M127 ð4rC Þ ≈ 1.31403 × 1026 m ≈ 13.89 Gly ≈ RU ð13.81 GlyÞ. ð30Þ
Thus, the special integer 137 (and hence the electric constant a) is related to
quantities that play a major role in microphysics, biophysics, or astrophysics. This
has two main consequences: (1) There might be a deep connection between the
various levels of complexity, as had been surmised by Dirac, Schrödinger,
Eddington, and others. (2) Following Pythagoras’ conjecture that ‘everything
proceeds from numbers’, constants occurring in these fields might be determined
not by some cosmic ‘natural selection’ or ‘anthropic principle’ [29], but by
numerical properties of specific numbers.
We have looked empirically for an expansion of the relative shift of a from 137
similar to that of ge from 2 given by Eq. (24). The following expansion:
which holds within 0.4 ppb, is surprisingly simpler and more accurate than
Eq. (24). A theoretical explanation is in progress.
One of the major problems in modern physics is the relation between electro-
magnetism, governed by quantum electrodynamics [7], and gravitation / inertia,
governed by general relativity [30]. In this paper, we shall recall two unconven-
tional attempts to shed some light on this relation.
The Dirac equation for the electron [10], which was derived in the frame of
special relativity theory, introduced four-component spinors expressing quantum
The Dirac Electron and Elementary Interactions … 371
properties. On the other hand, Einstein’s general relativity theory [30] expressed
space-time curvature, entailed by local equivalence of acceleration and gravity, in
the language of Riemann tensors. Among others, Chapman and Leiter [31] made
the Dirac equation conform to general relativity by expanding the principle of
covariance to spinor transformations.
For Mendel Sachs [32], the formal structure of quantum theory, including the
Pauli exclusion principle, can be derived from a low-energy, linear approximation
of a generally covariant, nonlinear field theory of inertia based on the basic ideas of
general relativity. On the other hand, John Macken [33] starts with the high-energy,
nonlinear vacuum fluctuations of quantized space-time and then derives the basic
ideas of general relativity, including the equivalence of acceleration and gravity, as
a low-energy, smooth limit.
Expressing the distance r between two identical particles (e, m0) as a multiple
N of their Compton diameter (reduced Compton wavelength): 2rC =−λ ≡ ℏ ̸ m0 c,
and scaling the electrostatic force: Fe = ke ⋅ e2/r2, and the gravitational force:
Fg = G ⋅ m20 /r2, to Planck units [34]: FP = c4 ̸G, EP = ðℏc5 ̸GÞ1 ̸2 , Macken [33] has
managed to express the two widely different forces as simply two different powers
of the rest mass energy of the two particles, E0 = m0c2:
F e = α E 20 ̸ N 2 , F g = E 40 ̸ N 2 , ð32Þ
a Fe ̸FP = Fg ̸a Fe = δ. ð33Þ
Here, the Compton diameter 2rC is replaced by the electric (quantum) force a Fe,
the larger Bohr radius a0 by the Planck force FP, and the smaller classical radius r0
by the gravific (relativistic) force Fg. According to Eq. (33), the gravific force is to
the electric force as the electric force is to the Planck force, the ratio of this relation
being a gravitational invariant: δ ≡ d − 1 , similar to the fine-structure constant:
α ≡ a − 1 , Eq. (26):
where m0 is the electron rest mass and mP is the Planck limit mass: mP = (ħ c/G)1/2.
Equation (33) is an indication that the two forces are deeply related through the
Compton diameter and then, to the particle spin, this diameter defining the
amplitude of Zitterbewegung, responsible for the spin properties (§ 2).
372 J. Maruani
In previous papers [1, 2], we showed that 2rC is also the geometric average of
the space-time curvatures, defined from general relativity [30], ‘inside’ the electron:
2rG = 2(G/c2) m0, and ‘outside’ a volume of radius rQ: 2RG = 2r2Q /rG. For rQ =
rC, one obtains a harmonic relation similar to Eqs. (27) and (33):
δ being defined in Eq. (34). Auxiliary relations resulting from Eqs. (27) and (35)
can be written:
From Eqs. (26, 37), canonical forms of the electric and gravific forces can be
drawn:
The value of dp in Eq. (37) is very close to the Catalan sum M’127 occurring in
Eq. (30) [28]. We have looked empirically for a series expansion of the relative
shift of dp from M’127 similar to that of a from M’7 = 137 or of ge from 2, given in
Eqs. (31) and (24), respectively. The following expansions:
′ ′
εp ≡ M127 − dp ̸ M127 ≈ 0.00478021
ð39aÞ
≈ ð1 ̸ 3Þð2 ̸137Þ − ð2 ̸5Þð2 ̸ 137Þ2 + ⋯ ≈ ðge ̸ aπÞ + 6ðge ̸ aπÞ2 − ⋯,
which hold within ∼0.7 ppm and ∼0.6 ppm respectively, are surprisingly simpler
and more accurate than one would expect from the poor accuracy of the measured
value of G used in Eq. (37). It can be noticed that 2/137 ≈ ge/a within <0.1% and
that ge/a ≈ 6−3 (6 being the first ‘perfect’ number) within <0.5%.
The Dirac Electron and Elementary Interactions … 373
If now we use the neutron (instead of the proton) mass mn in Eq. (37), we obtain:
dn ≈ 1.68862 × 1038, and a resulting εn ≈ 0.00751719 which could be expanded
as:
accurate to ∼ 5 ppm with the first 2 terms (∼ 2 ppm if 2/137 is replaced by ge/a)
and to ∼ 0.1 ppm with the 3 terms. This expansion looks more consistent than that
of Eq. (39a).
The gravitational and electromagnetic forces, which are the oldest known
(though rationalized only in modern times), thus share three features in common:
(1) they are long-distance and decrease as 1/r2 (Hooke-Newton and Coulomb laws);
(2) in Planck units, their magnitudes can be expressed as even powers of the
rest-mass energy (Eq. 32); (3) the parameters defining their strengths can be
expressed as Catalan integers of the Mersenne series: 2n − 1, with small shifts that
can be expanded, on the model of ge, as simple series involving a (or 137) and π
(Eqs. 24, 31, 39). However, these forces are governed by very different theories [7,
30], although various conciliation schemes have been attempted [e.g., 31–33].
Another force, of pure QED origin, is the Casimir force, which induces attraction
between two conducting plates separated by vacuum. It was proposed by Casimir in
1948 as a relativistically retarded van der Waals force. It has then been the subject
of various theoretical and experimental studies [35, 36], the first accurate mea-
surements being made not earlier than 1997. In the ideal case of perfect plates in
perfect vacuum at 0 °K, the force per unit area can be written as:
FC ̸ A = κ ℏ c ̸r 4 , κ = π 2 ̸ 240. ð40Þ
Contrary to those above, this force is very short ranged (<1 µm). However, three
interesting analogies can be drawn. (1) Similarly as empty space curvature (induced
by massive objects) is responsible for the gravity force [30], vacuum radiation
confinement (induced by conducting objects) is responsible for the Casimir force.
374 J. Maruani
0,0025
0,0015
0,0010
0,0005
0,0000
1 2 3 4 5 6 7 8 9 10 11
Distance (number N of Compton diameters λ)
Fig. 1 Reduced Casimir force: F C ̸E 20 (Eq. 41 with A ∼ π−λ2 ̸ 4Þ, versus reduced electric force:
F e ̸ E 20 (Eq. 32), as functions of the number N of Compton diameters between two electrons. The
curves cross after N = 2. The value N = 7 corresponds to ∼ 5% of a Bohr radius (Eq. 27)
This means that the reduced Casimir force F C exerted on a Compton-size square
(the size of an electron): A = −λ2 , has the same expression as F e but with the factor α
replaced by κ/N2. This factor can then be seen as a dimensionless parameter
characterizing the Casimir force. At a distance equal to a Compton diameter:
N = 1, the Casimir force is ∼ 6 times larger than the electric force. But it decreases
much faster with increasing N (Fig. 1). It may be worth investigating the role that a
Casimir attraction ‘inside’ the electron may play in its stability, as the strong force
in that of nuclei.
These forces, which are responsible for nuclear stability, are also very short ranged.
The strong nuclear force overcomes the electrostatic repulsion between protons,
while the weak nuclear force allows neutrinos to bind to nucleons [38].
The Dirac Electron and Elementary Interactions … 375
A number of model potentials have been proposed for the strong force (Yukawa
1934, Woods-Saxon 1954, Reid 1968, …), most of which are central scalar
potentials involving some exponential decrease. Nowadays, in the frame of the
standard model (where protons and neutrons are seen as made up of tight-bonded
quarks), the strong force appears as a residual, dispersion-like force. Its mean
magnitude, which depends on various factors, is known with poor accuracy [38], and
its dimensionless coupling constant: αs ≈ 0.1185 [39], is ∼ 16 times larger than the
fine-structure constant αe. Although this force does not follow a 1/r2 decrease, its
magnitude with respect to the electric force is consistent with the first sum M’3 of the
combinatorial hierarchy [28]: M’7/M’3 = 137/10 ∼ 14 [40].
The weak nuclear force was introduced in 1933 by Fermi to explain β decay
[38]. It has a number of specific features: (1) contrary to the other interactions, it
does not create bound states, but it allows transformation of a neutron to a proton or
vice versa by changing a quark flavor, this inducing β decay or electron capture;
(2) it is the only interaction that can violate parity symmetry; (3) contrary to the
other interactions, mediated by massless bosons, it is mediated by three very heavy,
short-lived bosons, two charged and one neutral. Besides, it is expected to decrease
with distance even faster than the strong nuclear force. However, in spite of these
crucial differences with the electromagnetic force, it has been unified with it in an
electroweak theory.
The weak force coupling constant is usually expressed in terms of a Fermi mass:
mF ≈ 5.730073 × 105 me ≈ 5.219743 × 10−25 kg, i.e.: GF ≡ ℏ3 ̸c m2F , which is in
J.m3 while ħ is in J.s [27]. One often uses the reduced constant: GF ̸ ðℏ cÞ3 =
1 ̸ m2F c4 ≈ 0.454380 × 1015 J − 2 [39]. But this is not dimensionless, as were
α and δ. One could also use the more familiar form: GF ̸ ℏ c = ðℏ ̸ mF cÞ2 ≈
0.454164 × 10 − 36 m2 , which is a Fermi-mass Compton-like area.
In order to get a dimensionless constant, some authors [29] have used the ratio
αw of this area to that for the electron mass (or its inverse aw). However, since the
weak force acts at the nucleon level, a more relevant choice would be to use the
ratio βw (or its inverse bw) involving the neutron mass:
The second ratio is ∼ 700 times smaller than the electric force constant α and
11,000 times smaller than the strong force constant αs , which is conform to
expected values.
Sanchez [27] has proposed the following relation (precise within 0.6 ppt)
between the gravitational, nucleoweak and electromagnetic force constants:
where the inverses of de, aw and ae are defined in Eqs. (34), (42) and (26),
respectively. It can be noted that the exponents in Eq. (43) are related to the
consecutive Catalan numbers 3, 7 and 127 by: 5 = (3 + 7)/2; 67 = (7 + 127)/2.
The question has often been raised as to whether the Newton and Coulomb forces
may not behave differently at very large or very short distances, or whether there
may not be other, undisclosed forces. Cosmological observations have led to the
conclusions that there is a dark energy and a dark matter constituting, respectively,
75% and 21% of the observable Universe. Could there also be forces still larger than
the electric and nuclear forces or smaller than the gravific force?
Looking back at the relation between Fe and Fg expressed by Eqs. (32, 33), one
may wonder whether the electric and gravific forces may not be just two of a series
of r−2-dependent forces acting at various levels, but with powers of E 0 other than 2
or 4. By extrapolation, one would then write:
F total = F a + Fe + Fg + F j + ⋯ = α2 + α E0 2 + E0 4 + α − 1 E0 6 + ⋯
ð44Þ
= α2 + α δp + δ2p + a − 1 δ3p + ⋯,
where the electric and gravific force terms are boldface. The divider N2, which is the
same for all r−2-dependent forces, has been omitted. Note that neither the Casimir
nor the nuclear forces, which are residual forces and decrease much faster than r−2,
fall in this frame. Each term in this series is derived from the previous one by
multiplication by α − 1 E 20 . These extra forces could then be expressed in terms of the
electric and gravific forces and, as a result of Eqs. (27) and (35), as ratios of particle
radii:
F a = F e 2 ̸F g , F j = F g 2 ̸F e . ð45Þ
The first force is 30 orders of magnitude larger than the strong nuclear force but,
as it is still 6 orders of magnitude smaller than the Planck force, it is not excluded
that it exists at a subnuclear level. The last force is 36 orders of magnitude smaller
than the gravity force and it appears much too small to show up in observable
effects, except maybe in the vicinity of neutron stars or black holes.
One may now turn back to the combinatorial hierarchy connection between
a and dp involved in Eqs. (31) and (39). Writing F x− 1 ≈ 2n − 1 (with 1 negligible
for n large) yields: na ≈ 14.18 ≈ 2 × 7+ (the increment to 7 being due to
137 > 127); ne ≈ 134.09 ≈ 127 + 7+; ng ≈ 253.99 ≈ 127 × 2, and nj ≈ 373.88
≈ 127 × 3 (from dp3) − 7+ (from a−1). However, none of the compound exponents
14 and 374 yields peculiar numbers comparable to those associated with a and dp.
7 Conclusions
In this paper, we first recalled the origins, features and main outcomes of the Dirac
equation: an underlying antimatter coupled with ordinary matter [9, 10], this
yielding a wave beat between the positive and negative energy states [15], resulting
in an internal motion at light speed within a Compton diameter [14]. Hence, the
spin kinetic momentum with quantum number s = ½ and magnetic moment with
gyromagnetic factor g0 = 2 [9, 10]. Various investigators had featured that the rest
mass would be related to the spin motion [16, 17], and we conjectured that the
electron can actually be seen as a massless charge spinning at light speed, the
observed rest mass stemming mainly from this very internal motion [1–4].
In this framework the Compton diameter−λ [12] plays a special role: as the range
of the internal motion, but also as the geometric average of the electron classical
radius r0 and the hydrogen Bohr radius a0, the ratio of this harmonic relation being
the fine-structure constant α = ke m20 ̸ℏc: −λ ̸ a0 = r0 ̸ −λ = α. It is also the geometric
average of the gravitational curvature diameters ‘inside’ and ‘outside’ the electron,
2rG and 2RG, the ratio of this harmonic relation being the gravitational invariant
2δe = 2 Gm20 ̸ ℏc: −λ ̸ 2RG = 2rG ̸−λ = 2δe .
Expressing the electric and gravific forces in Planck units, and the distance
between interacting particles as an integer number of Compton diameters, Macken
has shown [33] that they take the form of even powers of the rest mass energy. This
entails [2] a new harmonic relation, involving both α and δ, between the Planck,
electric and gravific forces: α − 1 Fe ̸ FP = Fg ̸ α − 1 Fe = δe . In this paper, we show that
a similar relation holds with the Casimir force but with α−1 replaced by a constant
378 J. Maruani
rapidly decreasing with distance. Other forces that would be expressed as similar
products of α and δp appear too large or too small to be observable.
Another link between the electric and gravific forces [27] results from the integer
parts of a and dp being the sums of Catalan numbers in the Mersenne series: 22 −
1 = 3, 23 − 1 = 7, 27 − 1 = 127, then M’7 = 3 + 7 + 127 = 137 ≈ a, and the
sum M’127 up to 2127 − 1 ≈ dp where dp ≡ δp− 1 and δp is defined as δe but with the
electron mass replaced by that of the proton (or the neutron). The magnitude of the
strong nuclear force, although it is a short distance one, seems consistent with the
sum M’3 = 10.
In this paper, we have discussed the deviations, from their integer approximates,
of the measured values of the free electron gyromagnetic factor ge, the fine-structure
constant inverse a, and the proton gravitational invariant inverse dp. After recalling
that the relative deviation of the measured ge from 2 (∼ 0.1%) was explained by
quantum field and size effects, we have tried to express that of a from 137 and that
of dp from M’127 as similar expansions.
Surprisingly, the relative deviation of measured a from 137 (< 0.03%) could
even better be expressed (with 0.4 ppb accuracy) as a simple 2-term series
involving even powers of π/137. Similarly, the relative deviation of measured dp
from M’127 (< 0.50%) could be expressed (with 0.6 ppm accuracy) as 2-term series
involving powers of either 2/137 ð ≈ ge ̸ aÞ or ge ̸ aπ ð ≈ 1 ̸ 63 Þ. For reasons yet
unexplained, the numbers ge, a, dp and π seem to be deeply related.
Acknowledgements I wish to thank the colleagues who helped me clarify these ideas at QSCP
meetings and elsewhere. Erkki Brändas, Uzi Kaldor, John Macken, Francis Sanchez, and Ivan
Todorov made especially useful comments. Thanks are due to my wife, Marja Rantanen, for
stimulating my speculations with inspiring piano playing.
References
1. Maruani J (2012) The Dirac electron: spin, Zitterbewegung, the Compton wavelength, and the
kinetic foundation of rest mass. Prog Theor Chem Phys B 26:23–46
2. Maruani J (2013) The Dirac electron as a massless charge spinning at light speed: implications
on some basic physical concepts. Prog Theor Chem Phys B 27:53–74
3. Maruani J (2015) The Dirac electron as a privileged road to the understanding of quantum
matter. Quantum Matter 4:3–11
4. Maruani J (2016) The Dirac electron: from quantum chemistry to holistic cosmology. J Chin
Chem Soc 63:33–48 and references therein
5. Thaller B (1992) The Dirac equation. Springer, Berlin
6. Sakurai J (1967) Advanced quantum mechanics. Addison-Wesley, Reading, MA, ch. 3
7. Feynman RP (1998) Quantum electrodynamics. Addison-Wesley, Reading, MA
8. Weinberg S (1995) The quantum theory of fields. Cambridge U P
9. Dirac PAM (1928) Quantum theory of the electron. Proc Roy Soc (London) A 117:610–624;
Quantised singularities in the electromagnetic field: ibid (1931) A 133:60–72; Theory of
electrons and positrons. Nobel lectures (1933) pp 320–325
10. Dirac PAM The principles of quantum mechanics. Clarendon Press, Oxford, 1st edn 1930,
4th edn 1958, chs 11–12
The Dirac Electron and Elementary Interactions … 379
11. de Broglie L (1924) Recherches sur la Théorie des Quanta. Thesis, Sorbonne, Paris; Ann Phys
10(III):22–128
12. Compton AH (1923) A quantum theory of the scattering of x-rays by light elements. Phys
Rev 21:483–502
13. Brandsden BH, Joachain CJ (1983) Physics of atoms and molecules. Longman, Harlow,
England, 1st edn, chs 1 and 5
14. Schrödinger E (1930) Über die kräftefreie Bewegung in der relativistischen Quantenmechanik:
Sitzungsber. Preuss Akad Wiss Berlin, Phys-Math Kl 24:418–428; Zur Quantendynamik des
Elecktrons: ibid (1931) 25:63–72
15. de Broglie L (1934) l’Electron Magnétique: Théorie de Dirac, Hermann, Paris, chs 9–22
16. Haisch B, Rueda A, Puthoff HE (1994) Inertia as a zero-point field Lorentz force. Phys Rev A
49:678–694 and references therein
17. Barut AO, Bracken AJ (1981) Zitterbewegung and the internal geometry of the electron. Phys
Rev D 23:2454–2463; D 24:3333–3342; Barut AO, Zanghi N (1984) Classical model of the
Dirac electron. Phys Rev Lett 52:2009–2012; Barut AO, Pavšič M (1987) Quantization of the
Zitterbewegung in the Schrödinger picture. Class Quant Grav 4:L131–L136
18. Todorov I (2015) Hyperlogarithms and periods in Feynman amplitudes. Opening lecture at
QSCP XX, Varna; published as CERN-TH-2016–042; and private communication
19. Dehmelt H (1993) Is the electron a composite particle? an experiment. Hyperfine Interact
81:1–3; updated and condensed version of the Nobel lecture in Les Prix Nobel 1989,
Stockholm, 1990, p 95
20. Maruani J (1980) Magnetic resonance and related techniques. In: Becker P (ed) Electron and
magnetization densities in molecules and crystals, NATO ASI series. Plenum, and references
therein
21. Sommerfeld A (1919) Atombau und Spektrallinien. Friedrich Vieweg, Braunschweig;
translated as Atomic structure and spectral lines. Methuen, London, 1923; and references therein
22. Eddington AS (1935) New pathways in science, Cambridge U P, ch. 11
23. Kaldor U, Eliav E, Landau A (2004) Relativistic electronic structure theory. Elsevier, ch. 2,
and private communication
24. Visscher L, Dyall KG (1997) At Data Nucl Data Tables 67:207
25. Kragh H (2003) Magic number—a partial history of the fine-structure constant. Arch Hist
Exact Sci 57:395–431
26. Gilson JG (1996) Calculating the fine-structure constant. Spec Sci Tech, 23 pp
27. Sanchez FM, Kotov VA, Bizouard C (2009) Evidence for a steady-state, holographic,
tachyonic, and supersymmetric cosmology. GED 20(3):43–53; Sanchez FM (2017) A
coherent resonant cosmology approach and its implications in microphysics and biophysics.
Prog Theor Chem Phys B 30:375–407, references therein, and private communication
28. Bastin T, Kilmister CW (1995) Combinatorial physics. World Scientific, Singapore
29. Carr BJ, Rees MJ (1979) The anthropic principle and the structure of the physical world.
Nature 278:605–612. See also: Barrow JD, Tipler FK (1986) The anthropic cosmological
principle. Oxford U P, and references therein
30. Einstein A (1916) Die Grundlage der allgemeinen Relativitätstheorie. Ann Phys (Leipzig)
49:769–823. See also: Eddington AS (1920) Space, time, and gravitation. Cambridge U P;
French edition completed with an updated mathematical presentation of general relativity.
Hermann, Paris, 1921. See also: Special issue commemorating Einstein A (2005). Ann Phys
(Leipzig) 14:1–204
31. Chapman TC, Leiter DJ (1976) On the generally covariant Dirac equation. Am J Phys
44:858–862
32. Sachs M (1986) Quantum mechanics from general relativity. Reidel, Dordrecht
33. Macken JA (2013) The universe is only spacetime. http://www.onlyspacetime/home; see also:
Spacetime-based foundation of quantum mechanics and general relativity. Prog Theor Chem
Phys A (2015) 29:219–245; and private communication
34. Pavšič M (2001) The landscape of theoretical physics: a global view. Kluwer, pp 347–352.
See also: http://en.wikipedia.org/wiki/planck_units
380 J. Maruani
35. Milton KA (2001) The Casimir effect: physical manifestations of zero-point energy. World
Scientific, Singapore
36. Genet C, Intravaia F, Lambrecht A, Reynaud S (2004) Electromagnetic vacuum fluctuations,
Casimir and van der Waals forces. Annales Fondation Louis de Broglie 29:311–328
37. André J-M, Jonnard Ph (2011) Effective mass of photons in a one-dimensional photonic
crystal. Phys Scr 84:035708 and references therein
38. Brown LM, Rechenberg H (1996) The origin of the concept of nuclear forces. Institute of
Physics Publishing, Bristol and Philadelphia
39. Beringer J et al (2012) Particle data group. Phys Rev D 86, 010001; for updates, see: http://
pdg.lbl.gov
40. It seems that Catalan numbers were known in the Antiquity. Not only the conspicuous 3 and 7
and their sum 10 (and various products and multiples), which occur in many traditions, but
also 127 (the number of provinces in Ahasuerus’ empire, cf Esther 1, 1) and the sum 137 (the
number of years Abraham’s elder son lived, cf Genesis 25, 17). The hypostyle room of
Ammon’s temple in Karnak, Egypt, has a total of 136 columns, the number that Eddington
initially proposed for the electric constant a [22]
A Simple Communication Hypothesis:
The Process of Evolution Reconsidered
Erkki J. Brändas
Dedicated to Arnold Trehub for his great visions and intuitions displayed in his Retinoid Model
of vision.
E. J. Brändas (✉)
Department of Chemistry, Uppsala University, Uppsala, Sweden
e-mail: [email protected]
1 Introductory Remarks
1
The ontic conceptualization is part of Heidegger’s neologisms [5]. His Being (Dasein) has a
pre-ontological-ontic signification, however, not to be understood as a biological human being.
This distinction is probably the reason why Heidegger takes a poetic turn as he investigates the
Question Concerning Technology and finds it to be but ‘a means to an end’.
A Simple Communication Hypothesis: The Process … 383
contributed his visions with a tongue in cheek, it should not give him absolute
immunity and exemption from liability. His attack on the claims of a philosophical
legislation for the exercise of science is stimulating and inspirational, yet it ignores the
foundation of abstracting science as an art of knowledge. To avoid the most bizarre
forms of philosophical relativism the scientific practice should focus on scientific
disciplines and their evolution rather than narrowing the perspective to a comparison
and review of various individual methodological rules and standards.
In the development of a scientific field, the historian of science, Mary Jo Nye [7]
identifies essentially six characteristics of a scientific discipline, i.e. genealogy, core
literature, practices, physical location, recognition and shared values. Within this
perspective it is clear that disciplines, as attractors for accumulated knowledge,
communicated between its members, focused on a particular sphere of Mother
Nature, do evolve with respect to styles and traditions compatible with existing
scientific paradigms and their inevitable changes. Hence scientific disciplines, as
examples of evolutionary lineages, imbedded in the human genetic code, do arise,
develop, evolve, die and diversify. For a detailed example of this process, see e.g.
the historical account of the emergence and evolution of quantum chemistry as a
sub-discipline and its importance in the province of science [8].
Recently the author advocated an evolutionary approach to represent and interpret
the origin and emergence of life, established by modern advances in chemical physics,
from unstable chemical states to biological evolution and order [9]. The Zero Energy
Universe Scenario (ZEUS) was elaborated commensurable with Darwin’s theory of
evolution, from the microscopic ranks to the cosmological domain.
In this contribution we will present an authentic stochastic communication
hypothesis, building on the results portrayed in [9], i.e. combining the novel formu-
lation of quantum mechanics for open systems, relativity theories, Gödel’s theorems,
Off-Diagonal Long-Range Order (ODLRO), the Correlated Dissipative Ensemble,
CDE, the Poisson distribution and associate encoding- decoding protocols for com-
munication between complex enough systems defining generic life forms. Some
examples of the various levels of organization, that abides by our Communication
Fig. 1 Communication
levels from the micro-to the COMMUNICATION
cosmic rank
function, homeostasis, information, ententional properties
Hypothesis, will be discussed in the final section, see also Fig. 1. We will also address
the longstanding disagreements between molecular- and evolutionary biologists [10]
on how natural selection acts, while supporting a trans-level semiotics that is
appropriate for geno-phenotypic translations including the role of somatic programs
in a Darwinian perspective. We will also compare and discuss our findings with other
results and strategies made recently in connection with the self-referential paradox
associated with the concept of consciousness. To make the presentation
self-contained, yet without bequeathing too many excruciating details, some brief
reviews are incorporated for the convenience of the prospective reader.
Today Darwin’s theory of evolution, in spite of initially being a badly split field, has
reached a certain consensus and acquired a synthetic unification known as the evo-
lutionary synthesis [10]. Nevertheless it is still classed as a sort of unfinished business
as the materialistic Neo-Darwinism2 has recently been seriously critiqued and
appraised [11]. Our answer to the issues and questions rendered above, has been
considered conceptually as the Paradigm of Evolution, see e.g. Refs. [9, 11] and
references therein. The formulation derives from a general extension of standard
mathematical representations of stationary states in separation from the environment
to so-called open (dissipative) systems.3 The logical framework should be general
enough to include the axioms of quantum mechanics. Below, we will, as already
mentioned, give a short account, but for more details we refer to [9, 11, 13].
Before activating a reading of the specifics needed for a more basic portrait, we
must recognize already at the start that there are many controversial undertakings
instigated by the various practitioners of science and their views of what should be
considered as fundamental concepts in the scientific description of Mother Nature.
For instance Barbour [14] rhetorically asks whether time is ‘real’, while at the same
time advocating a timeless theory of the universe. Furthermore Primas, in his Mind
and Matter article [15] projects a holistic reality, where time is not taken to be an a
priori concept. These examples of deviation from standard practice are well
enunciated, yet they run the risk of drawing unwanted consequences that incur more
problems than it solves, e.g. the unavoidable emergence of internal times, the
imperative dependence on boundary conditions, or paraphrasing Primas: ‘the death
of natural laws’ by being trapped in a Gödelian self-reference.4 We will return to
this question further below.
2
For discussions on the somewhat controversial concept of Neo-Darwinism, see e.g. [10, 11].
3
The concept of dissipative systems and their irreversible processes has been the focus of Pri-
gogine and the Brussels–Austin School, [12].
4
Note that the Gödel (self-referential) paradox can be translated to a consistently formulated
mathematical singularity of a suitably extended logic [16].
A Simple Communication Hypothesis: The Process … 385
Based on the argument just given above it follows that once space and time are
specified, momentum and energy are inevitable consequences of the declaration—
and vice versa. For instance, if some object, defined by its energy and momentum,
permits a portrayal within the basic deductive axioms of science as practised at
present in physics and formulated by the most succinct language we know today,
mathematics, we cannot avoid deductio in domum of the immaterial degrees of
freedom as given by time and space. Likewise, if we are able to differentiate some
non-stationary entities in space, we are in principle capable to discuss their
energy-mass-momentum relationships. The obvious conclusion is thus: inducing
the question whether time (or space) is not fundamental in some general setting
invites situations where the conjugate relationship viz. time–energy (or space–
momentum) would be relaxed. Although such situations might occur, e.g. in a black
hole, it might still be preferable to keep the deductive structure of a physical theory
intact, in order to investigate and analyse the emergence of any violations as they
386 E. J. Brändas
might appear. In what follows we will see by a simple realization how such situ-
ations arise under the most trivial conditions.
In particular we will analyse what happens when a non-zero rest-mass particle,
satisfying the Klein–Gordon (wave) equation, via some physical process (pair
annihilation etc.) produces photons. The interest in this question is motivated by
what happens to the degrees of freedom, i.e. the loss of the longitudinal degree
exhibited by massless particles (photons). The starting point will be a consideration
of the abstract Dirac kets in terms of the coordinate vector x⃗ and the linear
momentum p⃗
Here the energy is given by the mass relation E = mc2 , m the mass, c the velocity
of light and t, τ the time variable (operator).
The Klein–Gordon equation for a non-zero rest-mass particle is given in obvious
notation by
2
E02 2 Eop
− = p⃗ − ð3:1Þ
c2 c2
In order to analyse the intrinsic character of these two equations when the
non-zero rest-mass m0 goes to zero, we use Dirac’s trick to “take the square root of
the equation” by rewriting the observables in matrix form with Eqs. (3.1) and (3.2)
being their associated secular equation.
Hence our concern is the conjugate pair of observables usually represented in
operator form as (note that the time operator is trivially defined when the energy
interval is (–∞, +∞) as is also the time interval)
∂
Eop = iℏ ; p⃗op = − iℏ∇⃗x⃗ ð3:3Þ
∂t
and
A Simple Communication Hypothesis: The Process … 387
∂
τ = top = − iℏ ; x⃗op = iℏ∇⃗p⃗ ð3:4Þ
∂E
Hence we recognize Eq. (3.1) as the secular equation of the matrix defined in
Eq. (3.5) with the two eigenvalues ±λ, given by λ2 = m20 c2 , with m0 ≠ 0. Since the
formulation works for both variables and for operators, we will not explicitly single
out which reading is made unless the situation calls for a preference.
Similarly one obtains for the conjugate operators in (3.4) that Eq. (3.2) becomes
the secular equation of the (operator) matrix
cτ − ix⃗
ð3:6Þ
− ix⃗ − cτ
with ±cτ0 being the associated eigenvalues exhibiting the two time directions and
the left- and right-handed coordinate systems connected by space inversion sym-
metry. From (3.5) follows directly the usual approximations and estimates that
result in the conventional time-dependent Schrödinger equation as restricted to the
non-relativistic domain. With this inherent conjugate structure in mind, we remind
the readers of the attempts, see e.g. Refs. [14, 15, 21] to derive pioneering quantum
mechanics from general algebraic structures asserting that time (or space-time) is
not a primary concept. Yet the present description, as consistently built above, is
fundamental and commensurate with the deductive nature of the Lorentz
5
In fact a third comment is due, viz. the use of operators in the matrix calling for logical exten-
sions, see more in the next section. Note also the juxtaposition of the “arrow” of time and the parity
of space.
388 E. J. Brändas
transformation, see e.g. Refs. [18, 22], since from Eqs. (3.1) and (3.2) one obtains
straightforwardly the relativistic space-time scales of special relativity
m0
m = pffiffiffiffiffiffiffiffiffiffiffiffi ð3:7Þ
1 − β2
In this setting Eqs. (3.7) and (3.8) are valid irrespective of whether we are
representing classical wave propagation, quantum matter waves or classical
particles.
It is now possible to investigate the limit m0 → 0, observing the connection with
Maxwell’s equation for scalar and vector fields, by placing m0 = 0 in Eq. (3.1).
Something extraordinary happens with the character of the matrices of Eqs. (3.5)
and (3.6), when inserting the relation E = pc, with the momentum p⃗ assumed to be
directed along the x-axis. The result is a complex symmetric degenerate matrix
(reduced to the dimension of mass)
p ̸c − ip ̸ c 1 −i
= p ̸c ð3:5′ Þ
− ip ̸ c − p ̸c −i −1
In other words (3.5′) cannot be diagonalized since the two column vectors in the
matrix are linear dependent.
Analogously we obtain for (3.6) (since for the photon c2 = x⃗2 ̸ τ2 )
cτ − ix⃗ 1 −i 0 1
= cτ → 2cτ ð3:6′ Þ
− ix⃗ − cτ −i −1 0 0
We find to our surprise that that the complex symmetric matrices in Eqs. (3.5′)
and (3.6′) are nothing but a Jordan block of order 2, (a non-zero matrix, whose
square is zero) or, in more technical language, of Segrè characteristic 2. The linear
dependency leaves space-time with one less spatial dimension, i.e. along the x-axis.
Hence there is no longitudinal degree of freedom for a zero rest-mass particle with
6
The actual transformation is unitary, for details see e.g. Ref. [13].
A Simple Communication Hypothesis: The Process … 389
speed c, like the photon! As will be clear below, this implies that zero- and non-zero
rest-mass particles behave fundamentally different, which will become of crucial
importance in the case of the theory of general relativity to be reviewed below.
It is straightforward to extend our conjugate operator arrays above, to the general
case by the following modifications, see below, where μ is the gravitational radius,
G the gravitational constant, v = p ̸ c, r = jx⃗j, M a (usually large) spherically sym-
metric (non-rotating) mass,7 independent of m, i.e.
mð1 − κðr ÞÞ − iv
ð3:9Þ
− iv − mð1 − κðr ÞÞ
with8
mμ M
mκðr Þ = ; μ=G ⋅ 2 ð3:10Þ
r c
Note that as before, space-time and energy-momentum spaces associate the two
partitions into the material- and the immaterial sections of the Universe. Further-
more, as the area velocity multiplied by m is a constant of motion, one obtains for
local circular motion the boundary condition:
v = κ ðr Þc ð3:11Þ
7
Representing M by a ‘black hole’ shows the consistency between particle—antiparticle symmetry.
8
The matrix (3.9) has a direct link to Gödel’s self-referential paradox see e.g. [9].
9
Concepts like Nature, World, Universe, Cosmos, etc. are here used interchangeably.
390 E. J. Brändas
Leaving out the term r 2 dΩ2 the outcome becomes, not unexpectedly, the so-called
Schwarzschild10 line element ðm0 ≠ 0Þ:
with
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
A = B−1 = 1 − 2κðr Þ ð3:16Þ
∂ ∂
iℏ = Eop ðt Þ; top = − iℏ ð3:16Þ
∂t ∂Et
∂ ∂
iℏ = Eop ðsÞ; sop = − iℏ ð3:17Þ
∂s ∂Es
with
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ∂s pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Et = Es 1 − 2κ ðr Þ; = 1 − 2κðr Þ ð3:18Þ
∂t
10
As it is usually projected today.
11
In retrospect the singularity shares the same self-referential conundrum as we associate with
Gödel’s incompleteness theorem(s) [9, 16].
A Simple Communication Hypothesis: The Process … 391
where Es and Et represents the energy at the space time-points s and t respec-
tively. With the notation Es = ms c2 , one might identify the energy at a space-time
point s with the mass ms . One can see that the matrix Eq. (3.15) is commensurate
with the metric given by Eq. (3.12). For the case of photons (zero-rest-mass) one
obtains for κ ðr Þ ≠ 1 ̸ 2 that ds = 0 and
cAdτ − iBdx⃗ 1 −i 0 1
= cAdτ → 2cAdτ ð3:19Þ
− iBdx⃗ − cAdτ −i −1 0 0
which is consistent with Eq. (3.6′). One notes that the energy formula for
photons becomes
mð1 − 2κ ðr ÞÞ = p ̸ c
12
The common inconsistency in the vernacular is also known as the liar’s paradox and it has
recently been discussed under the name of The Pinocchio Paradox [23].
392 E. J. Brändas
dynamics of their conjugate partners. The strategy will be to develop the appro-
priate thermodynamics, invoking temperature and entropy to derive apt scales in
concert with appropriate ensembles and to find out how complex enough systems
interact or rather communicate with each other. Since we have presented the various
derivations in earlier contributions [9, 11, 16], we will for the most part only state
the mathematical results, as they are required as well as their physical interpretation.
Though our formulation has a rigorous origin in the axioms of quantum theory
as a trace algebra, see Ref. [20], in terms of general system operators, various
ensembles etc., there are important generalizations to quantum logic as engrained in
the illustrious theorem due to Gleason [24]. Hence the present formulation does not
only refer to pioneering quantum mechanical interpretations, but it also covers
interpretations that go beyond classical Boolean structures.13
A convenient starting point is the second order, reduced for N fermions, char-
acterized by the space-spin coordinates14 xk normalized to the number of pairings
(for details see [9, 11, 16])
0 0
Γð2Þ x1 , x2 jx1 , x2 =
Z 0 0 ð4:1Þ
N
Ψðx1 , x2 , x3 , . . . , xN ÞΨ* x1 , x2 , x3 , . . . , xN dx3 , . . . , dxN
2
with
n o
E = Tr H2 Γð2Þ ð4:2Þ
where the n-dimensional matrix γ (in principle specified below) is represented in the
space of the preferred basis jh⟩, the latter referring to appropriate pairs of light
carriers, like electrons described by paired orbitals denoted as geminals localized at
various nuclear centers. In fact the density matrix should describe, together with the
nuclear motion, the full dynamics of the system,15 although we have here only
13
For instance the demonstrated connection with Gödel’s incompleteness theorem contains a
non-Boolean probability that is extended to describing the interactions and communications
between Complex Enough Systems, CES. This prompts the system to be density matrixdenoted as
a Gödelian network [25].
14
We will denote the spatial coordinates with a vector notation, i.e. x⃗k .
15
Usually one invokes the Born-Oppenheimer approximation, i.e. separating the nuclear motion
from the many electron quantum problem. However, this will not be adequate here as will be seen
further below.
A Simple Communication Hypothesis: The Process … 393
Next we resume the exploration by confronting two major problems, i.e. dealing
with temperature dependences or accounting for the presence of quantum-thermal
correlations, and at the same time removing the notorious Born-Oppenheimer
approximation, in principle treating the nuclear degrees of freedom on an equal
footing with the electrons.
Let us treat the last problem first by exercising the so-called mirror theorem
employed by Carlson and Keller, [29], in connection with the reduced degrees of
freedom in connection with density matrix theory.17 In particular an application to
the entangled activities between the electronic motion and the movements of the
nuclei, implies that they are coupled through mirroring dynamics [31]. For instance
an electron orbital or geminal, hk projected around nucleus l, i.e. described locally
by the spatial coordinate x⃗l writes in Dirac notation
hk ðx⃗l Þ = ⟨x⃗l j hk ⟩
to be viewed as the kth electron orbital (geminal) as anticipated around the lth
nucleus, can also be interpreted as the scalar product between the lth nucleus
described by the Dirac ket,18 given by jx⃗l ⟩ and the electronic motion characterized
by jhk ⟩. Such scalar products suggest the key ingredients for the mapping between
the electronic and nuclear degrees of freedom. Hence system operators of the kind
Eq. (4.1) should in principle contain both electrons and nuclei, one of which,
electrons or nuclei, could be traced away in order to study the remaining dynamics
by suitable master equations, see e.g. Ref. [12]. The conclusion from the mirror
16
For fermionic systems there is also a 2n (n – 1) dimensional tail of unphysical pairings, which is
omitted.
17
The actual theorem goes back to Erhard Schmidt [30], the mathematician who was behind the
Gram-Schmidt orthogonalization.
18
The Dirac ket denotes a quantum state and its significance originates from the bra-c-ket form as a
scalar product.
394 E. J. Brändas
theorem is that the mappings between the electron- and the nuclear degrees of
freedom, and back, exhibits the same classical canonical form. This imparts the
possibility of a dual interpretation of the actual matrix representations of either the
electronic motion or the associated nuclear motion in principle bypassing the
Born-Oppenheimer approximation.
One may also view the entwined dynamics as a general scattering problem with
electronic particles scattered on a number of nuclear targets yielding consistent
scattering data, see e.g. Ref. [9]. This leads to the second problem of merging
quantum and thermal correlations at precise temperatures invoking associated time
scales and rigorous dissipative (open system) dynamics, which will be done in two
steps. First we need to match the relevant time scales of the system, via a proper
thermalization procedure obtaining the system operator for the open system at the
relevant temperature exhibiting authentic time scales for a realistic description of
non-equilibrium situations of relevance for, what we will denote, Complex Enough
Systems, CES, and their interpretation.
Consider an open system involving n bosonic or paired fermionic degrees of
freedom as an “incoming beam” impinging on a set of nuclear sites. The whole
scattering arrangement is characterized by a relaxation process19 with the time scale
τrel , assumed to be distinctly larger than the smaller thermal timescale τcorr = ℏ ̸kB T.
The protocol describes a process that one will, on the average, detect one quasi
particle degree of freedom in the differential solid-angle element dΩ during the
thermal timescale, e.g. with τcorr ≈2.46 × 10 − 14 s at 310 K. Straightforwardly one
obtains, with the incident flux, Ninc , being the number of particles/(degrees of
freedom) per unit area and time, Ns dΩ, the number of particles scattered into dΩ per
unit time being the standard relations between the differential- and the total cross
sections σ Ω and σ tot , the following formulas
Z Z
n dΩ Ns
Ninc = ; σ Ω dΩ = Ns dΩ = ; σ tot = σ Ω dΩ = dΩ ð4:5Þ
σ tot τrel τcorr Ninc
yielding the simple relationship between the two characteristic times, i.e.
n kB T τrel
= τrel = ð4:6Þ
4π ℏ τcorr
19
Generally speaking a specific molecular process is considered as a local perturbation out of the
quantum state which then relaxes back to thermodynamic equilibrium after a certain time τrel .
A Simple Communication Hypothesis: The Process … 395
Utilizing the standard Heisenberg relation between life times and energy widths
of the state, i.e. εk = ℏ ̸2τk one may express Eq. (4.6) as (with β = 1 ̸kB T, T the
absolute temperature and kB the Boltzmann constant)
2πðl − 1Þ
βεl = ð4:7Þ
n
Relations (4.5)–(4.7) display the relation between the time scales, the tempera-
ture T and the relevant dimension n of the non-equilibrium dissipative system.
The thermalization procedure, extended to the density matrix subject to the
Bloch equation with a Hamiltonian, H, producing thermal fluctuations commen-
surate with a given temperature, gives (note that the usual condition of the eigen-
state thermalization hypothesis is not fulfilled due to ODLRO)
∂ϱ 1
− = LB ϱ; LB ϱ = ðHϱ + ϱH Þ ð4:8Þ
∂β 2
which together with (4.7) yields the surprising result, with γ → γterm
with
0 1
0 λS 0 λL
B0 0 ⋯
B 0 0 CC
ωd = B ⋮ ⋱ ⋮ C = λL J n − 1 + λS J ð4:10Þ
@ 0 0 0 λS A
⋯
0 0 0 0
with J denoting the standard nilpotent matrix with zeros everywhere except with
ones above the diagonal. Note that n → N ̸2 imparts, in contrast to n → ∞ that
λL → λS → 1. In the basis f the density matrix ϱ in Eq. (4.10) writes, while here
leaving out the second term20 above, since it will
n hereoonly play a minor role in the
time evolution, n ≈ N ̸2, ρ∝ϱ normalized to Tr ρρ† = 1
1 1 1 n−1
ρ = pffiffiffiffiffiffiffiffiffiffi jh⟩Q⟨hj = pffiffiffiffiffiffiffiffiffiffi jf ⟩J⟨f j = pffiffiffiffiffiffiffiffiffiffi ∑ jfk ⟩⟨fk + 1 j ð4:11Þ
n−1 n−1 n − 1 k=1
with
20
When n → ∞, λL → N ̸2 will dominate ρ∝ N2 jf1 ⟩⟨fn j activating e.g. large-scale coherent axonal
firing.
396 E. J. Brändas
̸n
The unitary transformation B is finally given by21 with ω = eiπ
0 1
1 ω ω2 ... ωn − 1
1 B 1 ω3 ω6 ... ω3ðn − 1Þ C
B = pffiffiffi B C ð4:13Þ
n@⋮ ⋮ ⋮ ⋮ ⋮ A
1 ω2n − 1 ω2ð2n − 1Þ ... ω ðn − 1Þð2n − 1Þ
Lρ = ½Hρ − ρH
For instance applying L above to Eq. (4.11) gives to first order the sum of the
energy differences Ei − Ej . Since the thermalization, leading up to the successive
transitions in ρ, is brought about by the exchange of a thermal oscillation, due to the
energy super-operator in (4.8), the final outcome becomes the overall change
E1 − En , which is nothing but the thermal frequency associated with τcorr . Hence the
present Liouville picture describes the CDS by this frequency and with the char-
acteristic lifetime22 τrel .
As a result one obtains an entity, which we will call the Correlated Dissipative
Ensemble, CDE. The latter, by its construction, integrates a principal basis set of
21
This form was obtained in collaboration with C. E. Reid [32] see also [33].
22
A rigorous analytic continuation of L is given in [34]. While the real part of the eigenvalue
appear as energy differences, the imaginary parts add up here to be consistent with Eq. (4.6).
A Simple Communication Hypothesis: The Process … 397
CDS’s defined by Eq. (4.11), and denoted by jH ⟩, where each CDS is commen-
surate with the time scales τcorr and τrel related through Eqs. (4.6) and (4.7). In
terms of the frequency23 ω0 = ℏ ̸ kB T and τ = τrel each CDS is characterized by
n kT
= τrel = ω0 τ ð5:1Þ
4π ℏ
R
with Q = ω0 τdΩ = n serving as a quality factor for the CDE.24 Deriving the CDE
from CDS units, i.e. molecular aggregates, like the DNA and/or its protein over-
coats, in a biological system, will permit the definition of a particular cellular
quality value, Q. This is to some extent analogous to signal processing in com-
munication systems, where the quality aspects of cavity resonators, like musical
instruments etc., play a vital role in transmitting resonating qualities. For that reason
the quality value Q = n, for e.g. somatic cells in a multicellular organism, trans-
ferring vital details regarding their traits, provides cell recognition through the
process of “molecular communication”.
The CDE exhibits an analogical irreducible unit Q in the basis H of CDS base
units. A simplified analysis of the associated system operator shows that its diag-
onal elements can be determined by a general probability measure, let us say
pð AÞ = TrðρPð AÞÞ, (ρ a given density operator and P(A) a suitable self-adjoint
projection on Hilbert Space for the event A), and (1 – p(A))p(A) for any off-diagonal
one displaying ODLRO. An analogous analysis, cf. the CDS, leads through diag-
onalization and thermalization to a similar irreducible unit Q as in Eq. (4.11) but
generally with a different dimension m, in terms of H and the transformed F, cf. the
relations between h and f. The key difference is that τ = τrel play the role of the short
(relatively speaking) time scale with a longer scale emerging through Q see more
below. We can now write down the propagator corresponding as (with
I = ∑m k = 1 jFk ⟩⟨Fk j)
m−1
J = ∑ jFk ⟩⟨Fk + 1 j ð5:3Þ
k=1
jH ⟩ = jF ⟩B ð5:4Þ
The notation ω0 = ℏ ̸ kB T should not be confused with ω = eiπ ̸n defined in Eq. (4.13).
23
The Q-value should not be confused with the complex symmetric matrix Q.
24
398 E. J. Brändas
which in the classical case are related through the standard Fourier transform. In the
present case the extension requires the separation of positive and negative times
focusing our interest on the retarded propagator.25 Inserting the Liouvillian,
Eq. (5.2) in (5.5), one obtains
m − 1 t k
1 k
e − iP τ = e − iω0 t e − τ ∑
t t
J ð5:6Þ
k=0 τ k!
m
ðωτI − P Þ − 1 = ∑ ½ðω − ω0 Þτ + i − k ðiJÞðk − 1Þ ð5:7Þ
k=1
where one notes that the expansions are finite, limited by the dimension m. The
occurrence of higher order poles in Eq. (5.7) is reflected by the build up of a
polynomial in front of the decay factor in (5.6). As a result the usual microscopic
law of evolution dN ðt Þ = − ð1 ̸τÞN ðt Þdt modifies according to the highest power
m – 1 of J, see [13] for details
t
dN ðt Þ = t m − 2 m − 1 − N ðt Þdt; dN ðt Þ > 0; t < ðm − 1Þτ ð5:8Þ
τ
jh⟩ = jf ⟩B ð5:9Þ
or the CDS, relating n light carriers in a nuclear skeleton of n sites, treating the
nuclear and the electronic system on par, and
jH ⟩ = jF ⟩B ð5:10Þ
for the CDE, relating e.g. m cells in a certain organ localised at key positions in a
particular organ or organism. If m and n reveal no relation whatsoever, one might
not anticipate any affinity between the cells with non-commensurate Q-values.
25
A detailed exposition of the connection between the correlation function and the corresponding
spectrum, including analyticity requirement and appropriate integration contours have been dis-
cussed in the appendix of Ref. [35].
A Simple Communication Hypothesis: The Process … 399
However if they are the same—or as we will see containing many least common
multiples—they share the same time scale τ = τrel as well as displaying common
time factorizations as shown by the transformation B. Actually any perception,
depending on the rows of the matrix B, should amount to “translating” the cyclic
vectors of B via the transformation back to the “physical” basis h or H localized in
the cell and in the organism respectively. Hence one might envisage the translation
of compatible molecular information via the communication bearing transformation
B as authorized “messages” between the molecular-DNA-RNA-gene level and the
higher-level hierarchical organization leading to cellular function and biological
order.
On the condition that the Q-values in (5.9) and (5.10) are commensurate, it
suggests the possibility of a communication protocol given by the cyclic properties
and their nestings of the columns of B. For instance analysing the transformation in
Eq. (4.13), one observes the way the elements repeat themselves due to the
cyclicity of complex numbers. For instance choosing n = 12, computing the col-
umns of B, one may display the result as the simplified diagram below, in which
only the dimensions of the recurring vectors are indicated. Removing the first
column of one-dimensional units “1”, the resulting graph, containing 11 columns,
will look like
2
3 2 3
4 4
6 3 2 3 6
12 4 12 12 4 12 ð5:11Þ
6 3 2 3 6
4 2 4
3 3
2
where we do observe the obvious column symmetry of the table. From Eq. (5.6)
one finally realizes the typical behaviour of Poisson statistics that suggests a
communication concept that will share some analogy with the dynamics of a
telephone Call or Contact Center.26
26
This analogy has been carried out in some detail in relation to Trehub’s Retinoid Model of the
cognitive brain [36, 37]. The representation of neurons with a chemical synapse onto itself is of
particular significance for the Gödelian network [25].
400 E. J. Brändas
ascertain the Grand Master’s inner thoughts about the cornerstones of physics.
While Einstein did say that ‘time is an illusion’, Prigogine on the other hand
asserted that ‘we are the children of time’. As the prospective reader of this con-
tribution might have realized, energy-momentum and space-time are inevitably
coupled to each other as conjugate entities and that the present formulation, as a
consequence, invokes the characteristics of the Correlated Dissipative Ensemble,
CDE.
In previous studies we have emphasized the opportunity of cell recognition
suggested by the possibilities offered by the cellular Q-value prearranged by the
CDS as building blocks in every cell. This problem involves both inter- and
intra-cell communication. The lower- as well as the higher-level structures are
subject to Poisson statistics defining physical communication channels between and
within them. In particular we have focused on the CNS, the central nervous system,
in the presence of a special type of cells, i.e. neurons, with the dynamics charac-
terized by short time scale oscillations, building up pulses of light carriers (elec-
trons), i.e. spikes that correlate the basic dissipative systems e.g. cells/neurons, see
also footnote 21, and providing an irreducible coupling that arrange for the com-
munication between them. For more details regarding the usage of stochastic
backgrounds as objective physical communication channels, the former commen-
surate with the Poisson statistics, authorizing a direct call centre analogy, i.e. the
distribution of the rate of incoming phone calls received at each neuron, the
switchboard, see Refs. [9, 11, 25]. Note that the code for exchanging “communi-
cation” is given by the transformation B, Eq. (4.13), which applies in both limits:
N
λL → ; n→∞
2
and
N
λL → 1; λS → 1; n →
2
27
It is important to realize that this is not a “biological” symmetry, but induced by the perception of
mirror neuron images produced by the autapse.
28
This is a crucial feature, since according to the Charge, Parity, Time reversal, CPT, theorem,
believed to be the fundamental symmetry of physical laws, see also Ref. [38] the brain perceive
time reversal symmetry as a space parity inversion, which explains the Necker cube illusion [25].
A Simple Communication Hypothesis: The Process … 401
mirroring symmetry that is instigated by the feedback loops of the autapse. Hence
every trajectory in egocentric space exhibits a time-reversed copy in agreement with
the above-mentioned symmetry. This yields a simple understanding of the Necker
Cube Illusion [25] and the fact that our eyes see everything upside-down. The
examples given above relate to communication between cells connected to the
central nervous system, CNS. In what follows, we will concentrate on the other
question, the trans level communication.
As our conclusion concerns open questions in the philosophy of biology, see
Ayala and Arp [39] for recent debates, there appear some highly interesting work,
where the communication between the cells plays a preferential role [40]. In his
formulation of the First Principles of Physiology, integrating the self-organizing
status of the unicellular state, see also Ref. [41], John Torday, asserts that biological
organisms exist far from equilibrium, ‘circumventing’ the Second Law of Ther-
modynamics generating negative entropy within them, sustained by chemiosmosis.
Moreover he finds, going through the life cycles of biological phenomena, that
evolution on this level is deterministic, while at the same time arguing that the
unicell “finds itself” in an ambiguous situation with respect to its environment.29
Along these lines of reasoning, while acquiring epigenetic marks, one discerns
Lamarckian traits emphasizing environmental input over genetic evolution advancing
the so-called Epigenetic Inheritance Systems, EIS, [42], suggesting alternative possi-
bilities for extending the boundaries in evolutionary biology. For instance, recent
studies on genomic imprinting in mammals, see the recent review by Li and Sasaki [43],
is redolent of mechanisms relating to the reprogramming of pluripotent stem cells. One
should emphasize, however, that these models have no direct relation to quantum
chemical processes inside a cell. Nevertheless the general question remains, i.e. how
does selection act, what sort of entities are selected? To answer this question, one needs
to go to the intracellular level translating the intercellular signals on the level of the
DNA-protein coding to the higher-order level of cell-interactions. For instance, Torday
demonstrated how the parathyroid hormone-related protein PTHrP and leptin signalling
mechanisms, do facilitate a somatic program of lung evolution, offering ontogenic/
phylogenic links. In consideration of the present idea, it would be intriguing to find a
way to transcend the intracellular level, translating the molecular signals from the
microcirculation of the alveolus to alveolar metabolism.
One may enquire what time scale τ would be appropriate for the CDE in general.
While the CDS engender a relaxation time commensurate with its process forming
dynamics, it might be interesting to investigate e.g. the important role of the primary
cilium in modulating neurogenesis. They consist of micro-tubule-based organelles,
the latter also being part of the neuronal structure, which has been particularly
emphasized by Hameroff and Penrose [44] in connection with their “Orch OR” theory
for the appearance of consciousness in the universe.30 In eukaryotes the structure of
29
This is an intriguing declaration, since this is precisely “communication” formulated in this
article and programmed by Gödel-like sentences.
30
See also the discussions following their article in the Physics of Life Reviews.
402 E. J. Brändas
microtubules are long cylinders of α- and β-tubulin dimers,31 each with a molecular
weight of about 50 kDa formed in parallel association of thirteen protofilaments. In
total we have a weight of around 1000 kDa, corresponding to about 106 protons. With
a nuclear average of approximately 10 Da, one might expect roughly 105 nuclear
degrees of freedom to be at work in each of the ca 109 microtubules. Using Eq. (4.6)
one finds cilia beating frequencies of the order τrel ≈ 25 ms at human body temper-
atures, which appears to be in the right range for motile cilia mediated pathways. It has
also been noted that defects in the structure of the cellular primary cilium leads to
various disorders and impairments [45]. Hence looking at recurrent frequencies,
around 10–40 Hz, one will conclude that obstructing cell organelle movements might
severely disturb the channel for stochastic communication. Despite the view that
perhaps 10–20,000 neurons are involved in a perception instigated by specific spike
trains, the present CDS concept, nevertheless, implies a more complex understanding
in that a Gödelian network must have all neurons ‘on alert’ in case of need.
Another critical challenge, while retaining quantum chemistry applications to
biological systems, is the temperature dependence, the thermoregulation, and the
associated problem of the necessary prohibition of short-time32 decoherences. The
development and appearances of Correlated Dissipative Structures, CDS, not only
suggests answers to this dilemma, it also incorporates realistic time scales and the
linked nested ‘code-encode bearing’ transformation B. In this connection it is also
interesting to observe that so-called poikilotherms, i.e. organisms whose internal
temperatures varies considerably, have several different enzyme systems operating
at different temperatures, hence representable by different CDE’s.
Finally, a reminder regarding our biophysical organisation, termed a complex
enough system, CES, subject to the CDE, viz. the interactions between the cells
commensurate with the dynamics inside the cell, derived from a CDS, is neither
deterministic nor probabilistic.33 This incurs no contradiction, even if the biological
process may appear deterministic at the mesoscopic level of understanding.
The CDE is an irreducible ensemble defined in terms of the CDS, the latter
exhibiting an analogous irreducible structure. Within this framework one might
finally be reminded of the words of Ernst Mayr [10, 46] regarding the character-
istics of biological processes: a teleonomic process or behaviour is one that owes
its goal-directedness to the influence of an evolved program. Our interpretation
suggests that the stochastic hypothesis depicted here, i.e. a “Communication
31
Since the microtubules have a distinct polarity one obtains a direct coupling between the
mechanical- and the electromagnetic oscillations generating the train of neuronal spikes.
32
The notion ‘short’ is here in relation to any other time scales discussed in this article.
33
It has been stated that the mathematical structure of cosmos, with the origin of the physical laws
being in pure numbers, may discard the theory of biological evolution as well as being incom-
mensurate with the non-deterministic interpretation of quantum mechanics [47]. It is clear, how-
ever, from the present work that numerical structures, like the formula 1 + 1 = 2, reveals
important and necessary facts about our universe that are essential for the communication between
life forms and contingent on evolution, see e.g. a recent scientific analysis of ‘Science and Music’
[48] focusing on musical transcriptions in the living world.
A Simple Communication Hypothesis: The Process … 403
Simpliciter”, derived from first principles, can be understood and generated as such
an authentic process.
Acknowledgements I am grateful to the Chair of QSCP-XXI, Prof. Yan Alexander Wang, and
the Cochair Prof. Jean Maruani, for generously allowing me to present this work in the present
proceedings from the Vancouver meeting. This work has, over time, been supported by the
Swedish Natural Science Research Council, Swedish Foundation for Strategic Research, the
European Commission, and the Nobel Foundation.
References
22. Löwdin P-O (1998) Some comments on the foundations of physics. World Scientific,
Singapore
23. Eldridge-Smith P, Eldridge-Smith V (2010) The Pinocchio paradox. Analysis 70(2):212
24. Gleason AM (1957) Measures on the closed subspaces of a Hilbert space. J Math Mech 6:885
25. Brändas EJ (2015) ) Proposed explanation of the Phi phenomenon from a basic neural
viewpoint. Quant Biosyst 6(1):160
26. Yang CN (1962) Concept of off-diagonal long-range order and the quantum phases of liquid
helium and of superconductors. Rev Mod Phys 34:694
27. Sasaki F (1965) Eigenvalues of fermion density matrices. Phys Rev 138B:1338
28. Coleman AJ (1963) Structure of fermion density matrices. Rev Mod Phys 35:668
29. Carlson BC, Keller JM (1961) Eigenvalues of density matrices. Phys Rev 121:659
30. Schmidt E (1907) Math Ann 63:433
31. Brändas EJ, Hessmo B (1998) Indirect measurements and the mirror theorem. Lecture notes in
physics, vol 504, p 359
32. Reid CE, Brändas EJ (1989) On a theorem for complex symmetric matrices and its relevance
in the study of decay phenomena. Lecture notes in chemistry. Springer, Berlin, pp 325–475
33. Brändas EJ (2009) A theorem for complex symmetric matrices revisited. Int J Quant Chem
109:28960
34. Obcemea CH, Brändas EJ (1983) Analysis of Prigogine’s theory of subdynamics. Ann Phys
151:383
35. Brändas EJ (1997) Resonances and dilation analyticity in Liouville space. Adv Chem Phys
99:211
36. Trehub A (1991) The cognitive brain. MIT Press
37. Trehub A (2007) Space, self, and the theatre of consciousness. Conscious Cogn 16:310
38. Quack M (2011) Fundamental symmetries and symmetry violations from high resolution
spectroscopy. Handbook of high-resolution spectroscopy, vol 1, p 659
39. Ayala JF, Arp R (eds) (2010) Contemporary debates in philosophy of biology.
Wiley-Blackwell, Chichester, West Sussex
40. Torday JS, Miller WB Jr (2016) The unicellular state as a point source in a quantum biological
system. Biology 5(2):25
41. Torday JS (2016) Life is simple—biologic complexity is an epiphenomenon. Biology 5(2):17
42. Jablonka E, Lamb M (2005) Evolution in four dimension—genetic, epigenetic, behavioral,
and symbolic variation in the history of life. The MIT Press, Cambridge
43. Li Y, Sasaki H (2011) Genomic imprinting in mammals: its life cycle, molecular mechanisms
and reprogramming. Cell Res 21:466
44. Hameroff S, Penrose R (2014) Consciousness in the universe a review of the ‘Orch OR’
theory. Phys Life Rev Mar 11(1):39
45. Lee JH, Gleeson JG (2010) The role of primary cilia in neuronal function. Neurobiol Dis 38
(2):167
46. Mayr E (2004) What makes biology unique? Cambridge University Press, New York
47. Sanchez FM (2017) A coherent resonant cosmology approach and its implications in
microphysics and biophysics. In: Tadjer A, Pavlov R, Maruani J, Brändas EJ, Delgado-Barrio
(eds), Quantum systems in physics, chemistry, and biology, vol 30. Springer, Dordrecht,
p 379
48. Maruani J, Lefebvre R, Rantanen M (2003) Science and music: from the music of the depths
to the music of the spheres. In: Maruani J, Lefebvre R, Brändas EJ (eds), Advanced topics in
theoretical chemical physics and physics, vol 12. Kluwer, Dordrecht, p 479
Index
G P
General relativity, 370–372, 381, 385, 389 Parity violation, 96
Geometric phase, 265, 266, 268–270, 274, 275 Pionic atomic systems, 71, 72, 75
Gravitational invariant, 361, 371, 372, 377, 378 Planck units, 368, 371, 373, 377
Gyromagnetic factor, 365, 367, 369, 378 Polyborate, 108, 112, 120, 128, 144, 148
Population analysis, 112, 194
H Positron annihilation, 243–246, 248, 251–253,
Helium, 203, 204, 207–210, 253 255, 260
Homochirality, 95, 96 Proton transfer, 170, 179–182, 186, 188, 190,
Hydrogen, 30, 76, 82, 83, 100, 101, 108, 128, 191, 198, 200
130, 132, 135, 144, 151, 155–158, 164,
165, 169–171, 173, 190, 191, 193, 195, Q
198, 199, 201, 204, 206–210, 253, 255, Quantum chemistry, 3–8, 16, 72, 260, 383, 402
256, 265, 267, 268, 281, 289, 297, 305, Quantum electrodynamics, 362, 370
307, 309, 368, 369, 372, 377 Quantum theory of atoms in molecules, 46
Hydrogen bonding, 156, 164, 193, 201, 283,
309 R
Hylleraas-CI, 29–31, 33, 39 Radiation decay processes, 229
Hylleraas functions, 29, 30, 33 Relativistic energy approach, 57, 59, 66, 79,
231, 232
I Relativistic perturbation theory, 231, 234
Intramolecular hydrogen bond, 283, 305 Relativistic quantum chemistry, 72
Ionization potential, 65, 213, 224, 230, 247, Rhodium clusters, 213, 224
348 ROHF calculations, 223
J S
Jozimine A2, 305–307, 310–313, 316, 318, Semantic web, 3, 4, 7, 8, 18, 20, 26
320, 322, 324, 325 Solute-solvent interactions, 283
Space-time, 371, 372, 381, 387–389, 391, 400
K Spectroscopy of exotic atoms, 72
Kinetic-energy integrals, 29, 33, 39 Spin momentum, 98, 366
Stochastic communication, 382, 383, 402
L Structural indicator, 43, 44, 50, 53
Light velocity, 235, 365
T
M Transition metal, 215, 266
Magnetic moment, 88, 213, 222, 361, 364, 377 Trigger bond, 44–46, 50–53
Many-body theory, 243, 244, 259, 260 Trigger Bond Indicators (TBIs), 51, 52
Matter antimatter, 369
Molecular orbital, 119, 149, 204, 213, 215, 306 U
Multicharged ions, 55, 56, 66, 79 Ultracold chemistry, 265, 267
Multielectron atoms, 231 Ultracold collisions, 265
Mutual orientation, 283, 284, 288, 305, 308, Ultracold molecules, 267, 274
312, 314, 318
Index 407
V Web 3.0, 3
Vector spherical harmonics, 34, 35 Web of data, 3
Web of meaning, 3
W
Wave beat, 366, 377 Z
Weak bond, 43, 50–53 Zitterbewegung, 365–367, 371
Weakest bond, 43, 44, 50, 51, 53