Sei sulla pagina 1di 138

Theoretical and Computational Aspects of

Magnetic Molecules

Thesis

Submitted in partial fulfillment of the requirements for the degree of

Doctor of Philosophy

Submitted by

Md. Ehesan Ali


(02403001)

Under the Guidance of


Prof. S. N. Datta

Department of Chemistry
Indian Institute of Technology, Bombay
January, 2007
Dedicated to My Parents
Approval Sheet

Thesis entitled “Theoretical and Computational Aspects of Magnetic

Molecules” by Md. Ehesan Ali is approved for the degree of Doctor of

Philosophy

Examiners
_______________________
_______________________
_______________________

Supervisor
_______________________
_______________________
_______________________

Chairman
________________________
________________________
________________________

Date: ___________

Place: IIT–Bombay
Certificate of course work
This is to certify that Mr. Md. Ehesan Ali was admitted to the Ph.D. program on July 2002. He
successfully completed all the courses required for Ph.D. program. The details of the course work
are given below.

Sr. No. Course No. Course name Credit


1 CH 521 Interpretative Molecular Spectroscopy 6.00
2 CH 559 Solid State Chemistry and its Applications 6.00
3 CH 821 Topics in Chemistry I 6.00
Elements of Advanced Molecular
4 CH 842 6.00
Quantum Mechanics
5 CH 831 Advanced Laboratory Techniques 8.00
6 CHS 802 Seminar 4.00
Total credit 36.00

Place: IIT–Bombay Deputy Registrar (Academic)


Contents

Chapter 1: Magnetic Molecules

1.1. Introduction 1
1.2. Ferromagnetic Molecules 1
1.3. Insights from Literature Survey 2
1.4. Diradicals 3
1.4.1. Stable Diradicals 6
1.4.2. Triradicals and Polyradicals 7
1.5. Interesting Phenomenon related to Magnetic Molecules 8
1.5.1. Single Molecule Magnets 8
1.5.2. Photomagnetism 9
1.5.3. Spintronics 10
1.6. Scope of Molecular Magnetism 10
1.7. Objectives and Organization of the Thesis 11
1.8. References 12

Chapter 2: Theoretical Background

2.1. Introduction 17
2.2. Theoretical Background 19
2.2.1 Single Determinant Approach 19
2.2.2 Two-Determinant Configuration 21
2.2.3 Orbital Perturbation Theory 22
2.2.4 SCF energy 24
2.3. Spin Hamiltonian Treatment 26
2.3.1 Base line 27
2.3.2. Spin Hamiltonian 27
2.3.3. Expectation Values 29
2.4. Coupling Constants 30
2.5. Discussion 33
2.5.1. Factors Influencing Accuracy 33
2.5.2. Numerical Tests 33
2.6. Conclusions 36
2.7. References 37

Chapter 3: Organic Fused–Ring Diradicals

3.1. Introduction 40
3.2. Methodology 42
3.3. Energy Differences 43
3.3.1. 4-oxy-2-naphthalenyl methyl 44

i
3.3.2. 1,8-naphthalenediylbis(methyl) 45
3.3.3. 1-imino-1-naphthalenyl methyl 45
3.3.4. 1,8-naphthalenediylbis(amidogen) 46
3.3.5. 8-methyl-1-naphtyl carbine 47
3.3.6. 8-methyl-1-naphthalenyl imidogen 48
3.3.7. 8-methyl-1-naphthyl diazomethane 49
3.4. Conclusions 50
3.5. Reference 52

Chapter 4: Bis–Nitronyl Nitroxide Diradicals: Influence of Length and


Aromaticity of Couplers

4.1. Introduction 54
4.2. Theoretical Background 54
4.3. Computational Strategy 54
4.4. Results and Discussion 55
4.4.1. Rationalization 61
4.4.2. SOMO-SOMO Energy Level Splitting 63
4.4.3. Isotropic Hyperfine Coupling Constant 65
4.5. Diphenylene Acetylene Coupler 65
4.6. Conclusions 70
4.7. Reference 72

Chapter 5: Influence of Aromaticity in Intramolecular Magnetic Coupling

5.1. Introduction 74
5.2. Computational Methodology 74
5.3. Results and Discussion 76
5.3.1. Bond Order and Dihedral Angles 77
5.3.2. Nuclear Independent Chemical Shift 78
5.3.3. SOMO-SOMO Energy Splitting 81
5.3.4. Isotropic Hyperfine Coupling Constant 81
5.4. The m-Phenylene Couplers 82
5.4.1. Calculations 83
5.5. Conclusions 89
5.6. Reference 91

Chapter 6: Photomagnetism

6.1. Introduction 92
6.2. Technical Details and Results 93
6.3. Conclusion 98
6.4. References 98

ii
Chapter 7: Dinuclear Copper Complex

7.1. Introduction 99
7.2. Computational Methodology 102
7.3. Choice of Magnetic Orbitals 102
7.4. Results and Discussion 104
7.5. Conclusions 110
7.6. References 110

Conclusions 112
Summary i
Acknowledgment

iii
Chapter 1
Magnetic Molecules
This chapter describes a general introduction to magnetic molecules. A detailed literature
survey is also presented. The scope of molecular magnetism and the objectives of the Thesis
are discussed.
Chapter 1 Magnetic Molecules

1.1. Introduction
Magnetism has played a vital role in human civilization. The study of ferromagnetism
has been traditionally concerned exclusively with the study of transition elements (like Fe, Co
and Ni), alloys and metal oxides. This field of study has provided numerous technological
rewards based on the exploitation of such materials. In recent years, the focus of research on
magnetism has turned towards molecular systems and crystals. The reasons for this trend are
tunability of magnetic properties as a result of alterable chemical structures, bio-activity of
organic molecules, photo-control of chemical structure, and structure-property relationships.
In other words, the molecule can be tailored to exhibit desired magnetic properties.
Invariably, the building block of such magnetic materials are open-shell molecules
such as organic monoradicals, diradicals in which nonbonding molecular orbitals contain
unpaired electrons, and transition metal complexes with unpaired d-electrons. A large number
of high-spin magnetic molecules have been recently synthesized and investigated.1

1.2. Ferromagnetic Molecules


Molecular ferromagnetism results when the electronic spins in a single molecule are
coupled in a parallel orientation. In organic diradicals, two unpaired electrons at two different
non-bonded molecular orbitals (NBMOs) are coupled through the spacer in either parallel
(S=1) or anti-parallel (S=0) fashion, resulting in ferromagnetic or antiferromagnetic
interaction respectively. There are several rules to qualitatively predict the ground spin state
of such diradicals as discussed below.
According to Longuest-Higgins2, the number of NBMOs can be calculated as n = (N
− 2T) where N is the total number of the carbon atoms and T is the maximum number of
possible double bonds. This simple rule predicts the ground spin state of para-
benzoquinodimethan (1) (Figure 1.1). Four double bonds are possible and the number of
NBMOs is zero. Hence an antiferromagnetic interaction is observed. For meta-
benzoquinodimethan (2) (Figure 1.1), there are two NBMOs and according to the Hund’s
rule, a triplet (S=1) ground state is expected.3
From valence-bond formalism, Ovchinnikov4 suggested that the ground state spin S
can be determined as S = (n* −n)/2 for n* > n, where n* and n are the numbers of starred and
unstarred alternate carbon atoms. This rule predicts the meta- and para- benzoquinodimethan
to be in triplet and singlet ground spin states respectively (Figure 1.2).
1
Chapter 1 Magnetic Molecules

1 2
S=0 S=1

Figure 1.1. para- and meta- isomers of benzoquinodimethan.

* * * * *

* * * *

1 2

S=0 S=1

Figure 1.2. Predictions of ground state spin in starred/unstarred model and in the spin polarization
model.

The starred/unstarred model is very closely related to the concept of spin polarization
that can also predict the nature of the magnetic exchange interaction.5 Large positive spin
densities on atoms in conjugated systems induce small negative spin densities on neighboring
atoms. This spin polarization can be rationalized by considering the formation of the
chemical bond. The quantum chemical exchange favors a parallel orientation of the spins of
the electrons in σ- and π-orbital on the same atom over an antiparallel orientation.
Spin alternation rule that has been explained and demonestrated in UHF treatment can
also predict the ferro- and anti-ferromagnetic exchange interaction.6

1.3. Insights from Literature Survey


The first molecular ferromagnet, Fe(Cl)[S2CN(C2H5)2]2, was reported in 1972 and its
crystal was found to have ferromagnetic ordering at 2.43 K.7 Subsequently, Miller et al
synthesized a charge-transfer salt, composed a ferrocene derivative and tetracyanoethylene
with a Curie temperature of TC of 4.8 K.7
The first pure organic magnet, β crystalline phase of p-nitrophenyl nitronylnitroxide
(3) that orders magnetically at 0.65 K,8 is one of the major successes in modern research. In
1991, Wudl and coworkers discovered the organic molecular ferromagnet,
tetrakis(dimethylamino)ethylenefullerene[60] (TDAE-C60, 4) with Curie temperature of 16.1

2
Chapter 1 Magnetic Molecules

K.9 The origin of the ferromagnetism in TDAE-C60 has been the subject of various studies.
Its synthesis initially raised the hope that the higher TC values would soon be observed by
using other donors. But TDAE-C60 remained as the organic material with the highest TC until
1998 when Mihailovic reported 3-aminophenyl-methano-fullerene[60]-cobaltocene with a
slightly higher TC of 19 K.10 Rassat et al. have synthesized 1,3,5,7-tetramethyl-2-6-
diazaadamantane N, N'-dioxyl (5) (TC = 1.48 K).11 In 5, the two electron spins align in the
parallel fashion through intramolecular interaction, while the intermolecular interaction is
also ferromagnetic in nature. This phenomenon results a three-dimensional ferromagnetic
order below the Curie temperature.

O O
N O N
N
N O
O
Me2N NMe2 N
Me2N NMe2 O

3 4 5
Figure 1.3. Organic ferromagnetic molecules.

1.4. Diradicals
All the above examples of ferromagnetism arise from the intermolecular
ferromagnetic interactions. The intramolecular ferromagnetic order occurs in organic
diradicals and in dinuclear transition metal complexes. The simplest example of
intramolecular ferromagnetic interaction is trimethylenemethane (TMM, 6). TMM is a
diradical that has been widely investigated in different areas of chemistry. It was isolated in
matrix by Dowd in 1966.12 The significant and diverse impact of TMM has resulted from the
synthesis and studies of TMM derivatives that are stable and can be tailored to desired
molecular properties. TMM is the triplet ground state diradical with a very large singlet-
triplet energy gap (ΔEST). The estimated energy gap by photoelectron spectroscopy is in the
range of 13−16 Kcal mol−1.13 This range is confirmed by different computational studies.14
The large ΔEST indicates a very strong ferromagnetic interaction between the two radical
centers. As a result, TMM is a very attractive building block in molecular magnetism.

3
Chapter 1 Magnetic Molecules

Ar Ar
C C
Ar Ar

. .

TMM

Ar Ar
6 7

t-Bu t-Bu t-Bu O


O O N N
O t-Bu

t-Bu t-Bu

t-Bu t-Bu
+
O N
t-Bu O

8 9

Figures 1.4. TMM and its analogous diradicals.

Rajca et al. synthesized a new stable diradical 7, which is a 3-fold symmetric analogue
of TMM with no hetero atom substitution.14 This is similar but superior to 815 (Young’s
diradical) and 9.16 All these diradicals are stable and have triplet ground states. But only 7
can be extended to form polyradicals with a strong ferromagnetic coupling whereas the other
two are restricted due to their geometrical features. Recently, Shultz and coworkers have
synthesized stable diradicals 10-13 and studied their magnetic properties.17 These radicals are
TMM analogues. The authors noticed the structure-magnetic property relationship, that is, the
correlation of the exchange parameters with the phenyl-ring torsion angles (φ).
The strong ferromagnetic interaction can also be achieved via a benzene moiety (2).
The ground state of ortho- and para- substituted diradicals are singlet. Spectroscopic studies
of m-benzoquinodimethane (2) by Berson and their coworkers, Migirdicyan and Platz suggest
a triplet ground state18 and ab initio calculations predict a 10 Kcal mol−1 singlet-triplet energy
gap.19

4
Chapter 1 Magnetic Molecules

O O O O
N N
t-Bu N N
t-Bu
t-Bu t-Bu

t-Bu t-Bu

10 11

O O

N N O O
t-Bu t-Bu N N
t-Bu t-Bu

13
12

Figure 1.5. Recently synthesized TMM analog diradicals.

The exchange coupling constant (J) is consistent with ferromagnetic (JF) as well as
antiferromagnetic (JA) counterparts, that is, J = JF + JA. Antiferromagnetic coupling is
generally found to be more effective than the ferromagnetic interaction.20 Borden and
Davidson observed that the presence of non-disjoint MO (different orthogonal NBMO arising
from the common atom contributions) leads to a greater ferromagnetic coupling.21

14 15 16

Figure 1.6. Diradicals with two parallel couplers.

5
Chapter 1 Magnetic Molecules

The spectroscopic study of dimethylenecyclobutadiene (14), which is an example of


ferromagnetic coupling via two parallel coupling units, suggests a triplet ground state. The ab
initio calculations predict ΔEST = 17.7 kcal mol−1.22 The spectroscopic studies of the matrix-
isolated species 15 established its singlet ground state with ΔEST > −1 kcal mol−1.23 Ab initio
calculations suggest an energy gap of −5 kcal mol−1.24 For diradical 16, spin alternation rule
predicts the singlet ground state.25

1.4.1. Stable Diradicals


The first stable diradical (17) was synthesized by Schlenk almost a century ago, in
1915.26 It was identified as a ground state triplet by EPR measurement in matrix.27 A lot of
efforts has been applied to increase the stability and the magnetic exchange interaction in this
species.28 In fact, all the compounds 18-22 in Figure 1.7 show intense ESR spectra.29
The heteroatoms as spin sites can be attached to a strong ferromagnetic coupler. For
example, 10−13 are stable and ferromagnetically coupled diradicals. Stable, yet weakly
coupled diradicals are known.29−30 The stable diradicals with nitronyl nitroxide fragments are
fascinating species in modern research on molecular magnetism. A large portion of this thesis
will describe the computational studies on different Nitronyl Nitroxide systems. Till now a
large number of nitronyl nitroxide (NN) based diradicals has been experimentally
investigated.31
The intramolecular magnetic exchange coupling constant, as well as the
intermolecular interaction that depends upon the structure and the nature of a molecular
crystal, control the magnetic properties of a molecule-based magnetic material. An estimate
of the intramolecular exchange coupling constant is necessary prior to synthesizing a
successful magnetic material based on organic diradicals or transition metal complexes. The
recent development of computational techniques and theoretical methodologies has enabled
the prediction of magnetic properties of the precursors.32 Here we report the results of the
study of a series of nitronyl nitroxide based diradicals with different conjugated magnetic
couplers.
The magnetic couplings are generally found to arise from spin polarization and spin
delocalization.33 Lahti et al.5 investigated a large number of π-conjugated couplers. They
noticed that most of the spin density is localized on the two-singly occupied σ orbitals

6
Chapter 1 Magnetic Molecules

(SOMOs) centered on the radical atoms. The large spin population polarizes the π electrons
near the radical centre. The total π spin density sums to zero over all sites in the singlet state,
but the individual sites may be polarized to have positive or negative spin densities. The spin
polarization effect plays a major role in controlling the nature of the coupling. The presence
of non-bonding molecular orbitals (NBMOs) in organic diradicals makes it difficult to
properly evaluate the energy difference between the lowest states of different spin. The
expected ground state spin may be predicted either by molecular orbital (MO) calculation or
by a valence bond (VB) treatment. In the simple MO model, Hund’s multiplicity rules are
often applied to molecules having degenerate or nondegenerate NBMOs, with the prediction
of a triplet ground state. However, in a variety of conjugated systems the Hund’s criterion
does not necessarily follow, and a singlet ground state results. TME and its derivatives are
the simplest examples of such system. The low-spin nature of TME and the related disjoint
systems was explained by a VB-type electronic exchange. A number of derivations were
made to model the intramolecular exchange in connectivity-conjugated systems by
Ovchinikov,4 Klein,34 Borden and Davidson35 and Sinanŏglu.36 In all these cases the
simplistic MO theory and the Hund’s rule do not follow in a proper way. A large number of
computational studies have been performed on this issue.37−39 It is observed that the spin
polarization argument is more useful to understand the spin density distribution in an open
shell system.
R R t-Bu
t-Bu

t-Bu t-Bu

17 R=H, X=H 18
R=Me, X=H 19
R=i-Pr, X=H 20
R=Me, X=Me 21
R=CF3, X=H 22
Figure 1.7. The stable Schlenk diradicals (17) and its derivatives.

1.4.2. Triradicals and Polyradicals


Triradicals are relatively unusual. The nature of magnetic interaction between the
radical centers can be divided into three categories: (1) Two 1,3 connected benzene rings in a
7
Chapter 1 Magnetic Molecules

‘linear’ arrangement, (2) 1,3,5 substituted benzene ring, and (3) three 1,3,5 connected benzene
rings in a “closed loop” arrangement as shown in Figure 1.8. The expression for the magnetic
exchange coupling constant is J = ΔEQD, and 3J = ΔEQD, and 3J = ΔEQD for the three
categories respectively.

Δ EQD=J Δ EQD=3J Δ EQD=3J

Figure 1.8. Triradicals of three types.

Polyradicals (23) are most promising very-high-spin molecular systems. Rajca and
co-workers have synthesized decaradicals, which posses S = 5 high-spin state.40 Due to the
presence of spin defects in single ferromagnetic pathway, the experimental magnetization is
always less than the predicted value. To overcome this problem, Rajca introduced
calix(4)arene in the centre while triarylmethyl radicals are linked to it, and this provides
multiple pathways for the ferromagnetic interaction (24). 41

1.5. Interesting Phenomena Related to Magnetic molecules


Several interesting areas of research have been emerged from molecular magnetism.
Currently a large number of physicists and chemists are involved in these exciting areas. A
few of these are described below.

1.5.1. Single Molecule Magnet (SMM)


The single molecule magnet is an assembly of individual magnetic molecules. To be a
single-molecule magnet, the object must show a net magnetic spin and have no magnetic
interaction between molecules. Caneschi et al. reported the first magnetic molecules
[Mn12O12-(O2CMe)16(H2O)4] with a ground-state spin of S =10 in 1991.42 The SMM term
was first used by Hendrickson in 1996.43 The Mn12O12 complex and its analogous complexes
shows SMM behaviors but in the low temperature range with very small the spin barrier (∼50

8
Chapter 1 Magnetic Molecules

cm−1). To increase the critical temperature of SMM, (above which it behaves as

Ar
Ar Ar
Ar

Ar

n Ar Ar Ar

n n Ar
n
Ar
Ar
n
Ar
Ar Ar Ar
Ar

Ar Ar Ar
Ar

23 24

Figure 1.9. Magnetic plastics S=10 for 23 and S∼5000 for 24.

Ferromagnetic, antiferromagnetic or paramagnetic) the total spin quantum number S should


be very high and there must be a highly negative zero-field splitting parameters.

A number of theoreticians and experimentalists are making efforts to increase the


critical temperature of SMM to use them in molecular devices.44 Recently Soler et al. have
synthesized Mn12O12 based SMM.45 Davidson et al. calculated the magnetic properties of
these substances using local spin model.46

1.5.2. Photomagnetism
The extension of molecular photochromism results in the possibility of photo
switching of magnetic properties. Matsuda et al. have synthesized a large number of
photomagnetic molecules.47 The number of coordination compounds is limited to some
dithienylethylene derivatives,48 tetracyanoethylene organomettalic compounds, spin crossover
and valence-tautomeric complexes. From these perspectives, the Prussian blue analogue
complexes are the most promising. Recently Dei have synthesized a very interesting
9
Chapter 1 Magnetic Molecules

photomagnetic [{Cu(tren)-NC}6Mo(CN)2]8+ complex.49 A larger number of theoretical and


experimental work has been performed on this issue.50
1.5.3. Spintronics
Spintronics is the "spin-based electronics" and also known as magnetoelectronics.
This refers to the control of electric current through a manipulation of the spin of the
electrons. It has been extensively investigated using layers of ferromagnetic materials from
both fundamental and device application points of view, as in magnetic tunneling junctions
(MTJ).51-53

1.6. Scope of Molecular Magnetism

The materials which display cooperative magnetic phenomena yet are based on
molecular building blocks have added advantages over conventional magnets such as low
density, transparency, electrical insulation, low temperature fabrication as well as offering the
possibility of combining magnetic behavior (either cooperative, or isolated, as in “spin-
crossover compounds”) with other properties such as photo- or thermal responsiveness. In
addition, there are often useful processing advantages such as the ability to deposit the
materials as films, to functionalise them for attachment to substrates and possible
biocompatibility.

Magnetic properties of these materials can have important practical applications for
use in domestic appliances as well as in high-tech sciences. For example, in the rapidly
evolving world of micro-electronics, electronic circuits and storage devices are decreasing in
size and will eventually reach molecular dimensions. Thus, in future conventional semi-
conductor technology could be abandoned in favour of new materials with molecular
magnetic properties (“spintronics”) to build a computer. In addition to their small size,
magnetic materials consisting of molecular entities have the advantage that the molecular
precursors can be prepared under mild conditions in a directed synthesis and are therefore
easily integrated into materials with well-defined magnetic, magneto-optic or magneto-
electric properties. Spintronic devices are used in the field of mass-storage. Recently (in
2002) IBM scientists announced that they could compress massive amounts of data into a
small area, at approximately one trillion bits per square inch (1.5 Gbit/mm²) or roughly 1 TB

10
Chapter 1 Magnetic Molecules

on a single sided 3.5" diameter disc. Although some applications have already arisen from this
newly emerging field, it is still at an early stage of development. The ultimate rewards of
producing devices based on magnetic systems are increased processing power and
information storage, but this lie some way in the future. At present, the major requirements
for the further successful development of this highly promising field of materials science are
the production and investigation of a range of new systems and developing the basic research
into the underlying principles, namely a detailed description of molecular magnetism.

Molecular magnetism is a field of research where the investigation of the magnetic


properties of isolated molecules as well as of assemblies of molecules is undertaken. These
molecules may contain one or more magnetic centers. The assemblies of molecules occurring
in the solid state may be characterized by very weak interactions between the molecular
entities, thereby displaying magnetic behavior very similar to that of the isolated molecules.
They may consist of extended systems, in which strong magnetic interactions between the
molecular entities are responsible for bulk magnetic properties.

1.7. Objectives and Organization of the Thesis

The main aim of my PhD work has been to investigate the phenomenon called
molecular magnetism. A quantitative measure of magnetism in molecules is available in the
form of the intramolecular magnetic exchange coupling constant J. A knowledge of J helps
in predicting the magnetic moment and EPR frequency. Therefore, the objectives of this
thesis turn out as

1. To theoretically investigate J for diradicals and compare the derived formula with
those obtained by others;

2. To computationally investigate J for organic diradicals by using different ab initio


methodology;

3. To explore the possibility of photo-activation of magnetic properties;

4. To briefly investigate the characteristics of a transition metal complex diraical, and


understand the differences from organic diradicals.

The present thesis has been arranged according to the stated objectives. The arrangement is
as follows.
11
Chapter 1 Magnetic Molecules

A theoretical formalism for diradicals with nondegenerate HOMOs is given in


Chapter 2, where we present an N-electron interpretation of the spin Hamiltonian and
subsequently show that the Yamaguchi and GND formula arise from special types of
approximations.
In Chapter 3, fused-ring organic diradicals are investigated so as to determine the
ground spin states. The singlet-triplet energy gaps are calculated by using HF, post-HF and
DFT methodologies. A good correlation between the experimental and calculated results is
observed when the S−T energy difference is large. However, these molecules are less
important from the molecular magnetism point of view as these are very unstable.
The broken symmetry (BS) method of calculations has been adopted for the work
described in Chapter 4−7. Chapter 4 deals with a series of nitronyl nitroxide diradicals with
different linear and cyclic couplers. Effects of structural features, like chain length, dihedral
angles, etc., on J are investigated. We also predict the nature of magnetic exchange
interaction in a few new molecules.
The effect of aromaticity of benzene and polyacene couplers on the intramolecular J is
discussed in Chapter 5.
The investigation of the ground state magnetic properties of a few photochromic
molecules is described in Chapter 6. Some of these photochromic molecules can act as
photomagnetic switches.
The investigation of magnetic properties of a dinuclear copper complex is discussed in
Chapter 7.
Summary of the thesis and the conclusions are given subsequently.

1.8. References
1. (a) Gatteschi, D.; Khan, O.; Miller, J. S.; Palacio, F. (Eds.), Magnetic Molecular Materials, Kluwer
Academic Publishers, Dordrecht, 1991. (b) Kahn, O. (Ed.), Magnetism: A Supramolecular Function,
Kluwer Academic Publishers, Dordrecht, 1996. (c) Lahti, P. M. (Ed.) Magnetic Properties of Organic
Materials, Marcel Dekker, Inc., New York, 1999. (d) Miller, J. S.; Epstein.A. J. Angew. Chem., Int. Ed.
Engl. 1994, 33, 385. (e) Rajca, A. Chem. Rev. 1994, 94, 871. (f) Miller,J. S.; Drillon, M. (Eds.)
Magnetism: Molecules to Materials, Models and Experiments, Wiley-VCH, Weinheim, 2001. (g)
Miller, J. S.; Drillon, M. (Eds.) Magnetism: Molecules to Materials II,Molecule-Based Materials,
Wiley-VCH, Weinheim, 2001. (h) Miller, J. S.; Drillon, M. (Eds.) Magnetism: Molecules to Materials
III, Nanosized Magnetic Materials, Wiley-VCH, Weinheim, 2001. (i) Miller, J. S.; Drillon, M. (Eds.)

12
Chapter 1 Magnetic Molecules

Magnetism: Molecules to Materials IV, Molecule-based Materials (2), Wiley-VCH, Weinheim, 2002.
(j) van Meurs, P.J. High-Spin Molecules of p-Phenylenediamine Radical Cations, PhD Thesis,
Technische Universiteit Eindhoven, 2002. (k)
2. Longuet-Higgins, H. C. J. Chem. Phys. 1950, 18, 265.
3. Woodgate, G.K.; Elementary Atomic Structure, physics, McGraw-Hill, 1970.
4. Ovchinnikov, A. A. Theoret. Chim. Acta. 1978, 47, 297.
5. (a) Lahti, P. M.; Ichimura, A. S. J. Org. Chem. 1991, 56, 3030. (b) Ling, C.; Minato, M.; Lahti, P. M.;
van Willigen, H. J. Am. Chem. Soc.1992, 114, 9959. (c) Minato, M.; Lahti, P. M. J. Am. Chem. Soc.
1997, 119, 2187.
6. (a) Trindle, C.; Datta, S. N. Int. J. Quantum Chem. 1996, 57, 781. (b)Trindle, C.; Datta, S. N.; Mallik,
B. J. Am. Chem. Soc. 1997, 119, 12947.
7. (a) Wickman, H. H. J. Chem. Phys. 1972, 56, 976. (b) Chapps, G. D.; McCann, S. W.; Wickman, H. H.;
Sherwood, R. C. J. Chem. Phys. 1974, 60, 990. (c) DeFotis, G. C.; Palacio, F.; O’Conners, C. J.; Bhatia,
S. N.; Carlin, R. L. J. Am. Chem. Soc. 1977, 99, 8314.
8. (a) Awaga, K.; Maruyama, Y. Chem. Phys. Lett. 1989, 158, 556. (b) Awaga, K.; Maruyama, Y. J.
Chem. Phys. 1989, 91, 2743. (c) Awaga, K.; Inabe, T.; Nagashima, U.; Maruyama, J. J. Chem. Soc.,
Chem. Commun. 1989, 1617. (d) Awaga K.; Inabe, T.; Nagashima, U.; Maruyama, J. J. Chem.
Soc.,Chem. Commun. 1990, 520. (e) Turek, P.; Nozawa, K.; Shiomi, D.; Awaga, K.; Inabe, T.;
Maruyama, Y.; Kinoshita, M. Chem. Phys. Lett. 1991, 180, 327. (f) Kinoshita, M.; Turek, P.;Tamura,
M.; Nozawa, K.; Shiomo, D.; Nakazawa, Y.; Ishikawa, M.; Takahashi, K.; Awaga, K.; Inabe, T.;
Maruyama, Y. Chem. Lett. 1991, 1225. (g) Takahashi, M.; Turek, P.; Nakazawa, Y.; Tamura, M.;
Nozawa, K.; Shiomi, D.; Ishikawa, M.; Kinoshita, M. Phys. Rev. Lett. 1991, 67, 746. (h) Tamura, M.;
Nakazawa, Y.; Shiomi, D.; Nozawa, K.; Hosokoshi, Y.; Ishikawa, M.; Takahashi,M.; Kinoshita, M.
Chem. Phys. Lett. 1991, 186, 401.
9. (a) Allemand, P. M.; Khemani, K. C.; Koch, A.; Wudl, F.; Holczer, K.; Donavan, S.; Gruner, G.;
Thompson, J. D. Science 1991, 253, 301. (b) Narymbetov, B.; Omerzu, A.; Kabanov, V. V.; Tokumoto,
M.; Kobayashi, H.; Mihailovic, D. Nature 2000, 407, 883. (c) Mihailovic, D.; Arcon, D.; Venturini, P.;
Blinc, R.; Omerzu, A.; Cevc,P. Science 1995, 268, 400. (d) Tanaka, K.; Asai, Y.; Sato, T.; Kuga, T.;
Yamabe, T.; Tokumoto, M.Chem. Phys. Lett. 1996, 259, 574. (e) Mrzel, A.; Cevc, P.; Omerzu, A.;
Mihailovic, D. Phys. Rev. B 1996, 53, R2922. (f) Omerzu, A.; Mihailovic, D.; Tokumoto, M. Phys. Rev.
B 2000, 61, R11883.
10. Mrzel, A.; Omerzu, A.; Umek, P.; Mihailovic, D.; Jagličić, Z.; Trontelj, Z. Chem. Phys. Lett. 1998, 298,
329.
11. Chiarelli, R.; Novak, M. A.; Rassat, A.; Tholence, J. L. Nature 1993, 363, 147.
12. (a) Dowd, P. Acc. Chem. Res. 1970, 5, 242. (b) Dowd, P. J. Am. Chem. Soc. 1966, 88, 2587.
13. (a) Wenthold, P. G.; Hu, J.; Squires, R. R.; Lineberger, W. C. J. Am. Chem.Soc. 1996, 118, 475-476. (b)
Wenthold, P. G.; Hu, J.; Squires, R. R.; Lineberger, W. C. J. Am. Soc. Mass. Spectrom. 1999, 10, 800-
809. (d)Baseman, R. J.; Pratt, D. W.; Chow, M.; Dowd, P. J. Am. Chem. Soc. 1976, 98, 5726.

13
Chapter 1 Magnetic Molecules

14. (a) Ma, B.; Schaefer, H. F. Chem. Phys. 1996, 207, 31. (b) Cramer, C.J.; Smith, B. A. J. Phys. Chem.
1996, 100, 9664. (c) Li, J.; Worthington, S. E.; Cramer, C. J. J. Chem. Soc., Perkin Trans. 2 1998, 1045.
15. Rajca, A.; Utamapanya, S. J. Org. Chem. 1992, 57, 1760-1767.
16. (a) Mukai, K.; Ishizu,K.; Kakahara, M.; Deguchi, Y. Bull. Chem. Soc. Jpn. 1980, 53, 3363. (b)
Willigen, H.; Kirste, B.; Kurreck, H.; Plato, M. Tetrahedron 1982, 38, 759. (c) Bock, H.; John, A.;
Havlas, Z.; Bats, J. W. Angew. Chem., Int. Ed. Engl. 1993, 32, 416.
17. (a) Itoh, T.;Matsuda, K.; Iwamura, H.; Hori, K. J. Am. Chem. Soc. 2000, 122, 2567. (b) Oniciu, D. C.;
Matsuda, K.; Iwamura, H. J. Chem. Soc., Perkin 2 1996, 907.
18. (a) Rule, M.; Math, A. R.; Seeger, D. E.; Hilinski, E. F.; Dougherty, D. A.; Berson, J. A. Tetrahedron
1982,38,787. (b) Goodman, J. L.; Berson, J. A. J. Am. Chem. Soc. 1985, 107, 5409. (c) Wright, B. B.;
Platz, M. J. Am. Chem. Soc. 1983, 105, 628. (d) Migirdicyan, E.; Baudet, J. J. Am. Chem. Soc. 1975, 97,
7400.
19. (a) Fort, R. C., Jr.; Getty, S. J.; Hrovat, D. A.; Lahti,P. M.; Borden, W. T. J. Am. Chem. Soc. 1992, 114,
7549. (b) Kato, S.; Morokuma, K.; Feller, D.; Davidson, E. R.; Borden, W. T. J. Am. Chem. Soc. 1983,
105, 1791.
20. Girerd, J.J.; Journaux, Y.; Khan, O. Chem. Phys. Lett. 1981, 82, 534.
21. Borden, W. T.; Davidson, E. R. J. Am. Chem. Soc. 1977, 99, 4587.
22. (a ) Snyder, G. J.; Dougherty, D. A. J. Am. Chem. Soc. 1985, 107, 1774. (b) Dowd, P.; Paik, Y.H. J.
Am. Chem. Soc. 1986, 108, 2788. (c) Du, P.; Hrovat, D. A.; Borden, W. T. J. Am. Chem. Soc. 1989,
111, 3773. (d) Hudson, B. S.; Ziegler, L. D. In Excited States; Lim, E. C., Ed.; Academic Press: New
York, 1982, Vol. 5, p 70-74, 118-119.
23. Reynolds,J. H.; Berson, J. A.; Kumashiro, K. K.; Duchamp, J. C.; Zilm, K. W.; Rubello, A.; Vogel, P. J.
Am. Chem. Soc. 1992, 114, 763.
24. Reynolds, J. H.; Berson, J. A.; Scaiano, J. C.; Berinstain, A. B. J.Am. Chem. Soc. 1992, 114, 5866.
25. Wang, Di-Fei; Wu, Yun-Dong, J.Theo. Comp. Chem. 2004, 3, 51.
26. Schlenk, W.; Brauns, M. Chem. Ber. 1915, 48, 661,716.
27. Sholle, V. D.; Rozantaev, E. G. Russ. Chem. Reo. 1973, 42, 1011.
28. (a) Rajca, A.; Utamapanya, S.; X u, J. J. Am. Chem. Soc. 1991,113, 9235. (b)Veciana. J.; Rovira. C.;
CresDo. M. I.; Armet. O.; Dominno. V. M.; II Palacio., F. J. Am. Chem. Soc. 1991, 113, 2552.
29. Shultz, D.A.; Fico, R.M.; Lee, H.; Kampf, J.W.; Kirschbaum, K.; Pinkerton, A.A.; Boyle, P.D J. Am.
Chem. Soc. 2003, 125, 15426.
30. Shultz, D. A.; Fico, R. M., Jr.; Bodnar, S. H.; Kumar, R. K.; Vostrikova, K. E.; Kampf, J. W.; Boyle, P.
D. J. Am. Chem. Soc. 2003, 125, 11761.
31. Takui, T.; Sato, K.; Shiomi, D.; Ito, K.; Nishizawa, M.; Itoh, K. Syn. Metals 1999, 103, 2271. (b)
Romero, F. M.; Ziessel, R.; Bonnet, M.; Pontillon, Y.; Ressouche, E.; Schweizer, J.; Delley, B.; Grand,
A.; Paulsen, C. J. Am. Chem. Soc. 2000, 122, 1298. (c) Nagashima, H.; Irisawa, M.; Yoshioka, N.;
Inoue, H. Mol. Cryst. Liq. Cryst. Sci. Technol. Sect. A: 2002, 376, 371. (d) Rajadurai, C.; Ivanova, A.;
Enkelmann, V.; Baumgarten, M. J. Org. Chem. 2003, 68, 9907. (e) Wautelet, P.; Le Moigne, J.;

14
Chapter 1 Magnetic Molecules

Videva, V.; Turek, P. J. Org. Chem. 2003, 68, 8025. (f) Deumal, M.; Robb, M. A.; Novoa, J. J.
Polyhedron 2003, 22(14-17), 1935.
32. (a) Kahn, O. Molecular Magnetism; VCH: New York, 1993. (b) Goodenough J. B. Magnetism and the
Chemical Bond; Interscience: New York, 1963. (c) Coronado, E.; Delhaè, P.; Gatteschi, D.; Miller, J. S.
Molecular Magnetism: From Molecular Assemblies to the Devices, Eds.; Nato ASI Series E, Applied
Sciences, Kluwer Academic Publisher: Dordrecht, Netherland, 1996; Vol. 321 (d) Benelli, C.;
Gatteschi, D. Chem. Rev. 2002, 102, 2369. (e) McConnel, H. M. J. Chem. Phys. 1958, 28, 1188.
33. Dietz, F.; Tyutyulkov, N. Chem. Phys. 2001, 264, 37.
34. (a) Klein, D. J. Pure Appl. Chem. 1983, 55, 299. (b) Klein, D. J.; Alexander, S. A. In Graph Theory and
Topology in Chemistry; King, R. B., Rouvay, D. H., Eds.; Elsevier: Amsterdam, The Netherlands,
1987; vol. 51, p 404.
35. Davidson, E. R.; Clark, A. E. J. Phys. Chem.A . 2002, 106, 7456.
36. Shen, M.; Sinanŏglu, O. In Graph Theory and Topology in Chemistry; King, R. B., Rouvay, D. H.,
Eds.; Elsevier: Amsterdam, The Netherlands, 1987; vol. 51, p 373.
37. (a) Nachtigall, P.; Jordan, K. D. J. Am. Chem. Soc. 1992, 114, 4743. (b) Nachtigall, P.; Jordan, K. D. J.
Am. Chem. Soc. 1993, 115, 270.
38. Rijkenberg, R. A.; Buma, W. J.; van Walree, C. A.; Jenneskens, L. W. J. Phys. Chem. A 2002, 106,
5249.
39. Datta, S. N.; Mukherjee, P.; Jha, P. P. J. Phys. Chem. A 2003, 107, 5049.
40. Rajca, A.; Utamapanya, S.; Thayumanavan, S. J. Am. Chem. Soc. 1992, 114, 1884.
41. Rajca, A.; Wongsriratanakul, J.; Rajca S. Science 2001, 294, 1503.
42. Caneschi, A.; Gatteschi, D.; Sessoli, R.; Barra, A. L.; Brunel, L. C.; Guillot, M. J. Am. Chem. Soc.
1991, 113, 5873.
43. Aubin, S.M.J.; Wemple, M.W.; Adams, D.M.; Tsai, H.-L.; Christou, G.; Hendrickson, D.N. J. Am.
Chem. Soc. 1996, 118, 7746.
44. (a) Wernsdorfer, W.; Sessoli, R. Science 1999, 284, 133. (b) Leuenberger, M. N.; Loss, D.; Nature
2001, 410, 789.
45. Soler, M.; Rumberger, E.; Folting, K.; Hendrickson, D. N.; Christou, G.; Polyhedron 2001, 20, 1365.
46. Davidson, E.R.; Clark, A.E.; Mol. Phys. 2001, 100, 373.
47. (a)Tanifuji, N.; Matsuda, K.; Irie, M. Polyhedron 2005, 24, 2484. (b) Matsuda, K.; Irie, M. Polyhedron
2005, 24, 2477. (c) Matsuda, K. Bull. Chem. Soc. Jap. 2005, 78, 383. (d) Tanifuji, N.; Irie, M.;
Matsuda, K. J. Am. Chem. Soc. 2005, 127, 13344. (e) Tanifuji, N.; Matsuda, K.; Irie, M. Org. Lett.
2005, 7, 3777. (f) Matsuda, K.; Irie, M. J. Photochem. Photobio. C: Photochem. Rev. 2004, 5, 69. (g)
Matsuda, K.; Matsuo, M.; Irie, M. J. Org. Chem. 2001, 66, 8799.
48. Gütlich, P., Garcia, Y., Woike, T. Coord. Chem. Rev., 2001, 219, 839.
49. Dei. A. Angewandte Chem. 2005, 44, 1160.
50. Sato, O.; Iyoda, T.; Fujishima, A.; Hashimoto, K. Science 1996, 272,704. (b) Sato, O.; Einaga, Y.;
Iyoda, T.; Fujishima, A.; Hashimoto, K. J. Electrochem. Soc. 1997, 144, L11. (c) Sato, O.; Einaga, Y.;
15
Chapter 1 Magnetic Molecules

Iyoda, T.; Fujishima, A.; Hashimoto, K. J. Phys. Chem. B 1997, 101, 3903. (d) Einaga, Y.; Ohkoshi, S.-
I.; Sato, O.; Fujishima, A.; Hashimoto, K. Chem. Lett. 1998, 585. (e) Sato, O.; Einaga, Y.; Fujishima,
A.; Hashimoto, K. Inorg. Chem. 1999, 38, 4405. (f) Verdaguer, M. Science 1996 , 272, 698. (g) Raghu
athan, R.; Ramasesha, S.; Mathoniere, C.; Marvaud, V. Phys. Rev. B 2006, 73, 045131.
51. Johnson, M.; Silsbee, R. H. Phys. ReV. Lett. 1985, 55, 1790.
52. Wolf, S. A.; Awschalom, D. D.; Buhrman, R. A.; Daughton, J. M.; von Molna´r, S.; Roukes, M. L.;
Chtchelkanova, A. Y.; Treger, D.M. Science 2001, 294, 1488.
53. Awschalom, D. D.; Samarth, N.; Loss, D. Semiconductor Spintronics and Quantum Computation;
Springer: Berlin, 2002.

16
Chapter 2
Theoretical Background
This chapter describes the theoretical background of the determination of magnetic
exchange coupling constant from first principle calculations. An N-electron spin
Hamiltonian is formulated for diradicals having non-degenerate highest occupied molecular
orbitals. At first, energy expressions are obtained for singlet, broken-symmetry and triplet
single-determinant wave functions of unrestricted Hartree-Fock treatment. Total energy
values for the two-determinant singlet and triplet configurations that can be obtained from a
self-consistent-field treatment are determined next by using the orbital perturbation theory.
This leads to an energy ordering, which is expected to be valid also in an unrestricted
Hartree-Fock Kohn-Sham treatment. The spin Hamiltonian is based on this ordering.
Using the spin Hamiltonian, we obtain an expression for the energy differences from which
the Yamaguchi and GND formula for J can be easily obtained.
Chapter 2 Theoretical Background

2.1. Introduction
The interaction between two magnetic sites A and B in a diradical species is usually
expressed by the Heisenberg two-spin Hamiltonian
G G
H=E0 − 2 JSA ⋅ SB (2.1)
G G
where SA and SB are the respective spin angular momentum operators for the monoradical

fragments. A positive sign of J indicates a ferromagnetic interaction, whereas a negative sign


indicates an antiferromagnetic interaction. The eigenfunctions of the Heisenberg Hamiltonian
are eigenfunctions of S2 and Sz where S is the total spin angular momentum, and J is directly
related to the energy difference between the eigenstates. For a diradical,
E(S=1 ) − E(S=0 )= − 2 J . The two-spin description of magnetic interaction has been
extensively correlated with the electronic structure of diradical systems.1 Recently a large
number ab initio calculations have been performed to evaluate J.2 A proper mapping between
the Heisenberg spin eigenstates and the electronic states is necessitated for the above
procedure. This is true in principle but computationally very expensive.
An alternative approach has been proposed by Noodleman so as to reliably compute
the magnetic exchange coupling constant by density functional theory with less computational
effort.3 The spin polarized, unrestricted formalism and a broken-symmetry (BS) solution for
the lowest spin-state are required in this method. The BS state is not an eigenstate of Ĥ . It is
an equal mixture of singlet and triplet states. The coupling constant can be written as
( EBS − ET′ )
J= (2.2)
1 + Sab 2

where Sab is the overlap integral between the spatial part of α and β orbitals in the BS

solution. Eq. (2.2) is valid for the S=1/2 interacting spins. The quantity EBS is the energy of
the BS solution and ET ′ is the energy of the triplet state in the unrestricted formalism using

the BS orbitals. In a single-determinant approach, ET ′ can be approximated by the energy of

the true triplet state ( ET ′ ≈ ET ) because of the very less spin contamination in the high spin

state.
It is observed in literature that Eq. (2.2) is used in the strongly localized or orthogonal
limit where Sab → 0 as well as the strongly delocalized limit where Sab → 1. The current

literature is full of controversy regarding the choice of limit. Generally, in density functional
17
Chapter 2 Theoretical Background

(DFT) based calculations the magnetic orbitals are more delocalized than those obtained from
the unrestricted Hartree-Fock (UHF) calculations, and some of the authors have
recommended the use of Eq. (2.2) also in the strongly delocalized limit.4 However, Bencini et
al. have argued that Eq. (2.2) is to be restricted to the strongly localized limit.5 It has also
been concluded that the limit should be chosen on the basis of the proximity of calculated and
experimental values, rather than a consideration of rigorous theoretical complications.6
Illas et al.7 have shown that the most often-quoted trend concerning the much larger
degree of delocalization of magnetic orbitals obtained from DFT as opposed to UHF is not
fully justified. They have recommended the use of the strongly localized limit for the general
cases. In the strongly delocalized limit, Equation (2.2) becomes
2J ≈ EBS − ET . (2.3)

In this situation singlet becomes degenerate with the broken-symmetry state which does not
have any scientific evidence. Despite these problems and several other deficiencies in DFT as
recently mentioned by several authors, Eq. (2.3) produced very impressive numerical results
for some systems by using the so-called B3LYP exchange correlation functional treatment.9
A large literature is found on this issue.10
The following spin projected equations are commonly used in the investigation of J in
different circumstances:
(i) the GND equation11,12

( DFT E BS − DFT E T )
J (1) = (2.4)
Smax 2
(ii) the Bencini-Ruiz equation13,14

(2) ( DFT E LS − DFT ET )


J = (2.5)
Smax ( Smax + 1)
(iii) the Yamaguchi equation15

( DFT E BS − DFT E T )
J (3) = . (2.6)
< S 2 >T − < S 2 > BS
To settle the controversy in the choice of the correct expression for J, Neese has recently
analyzed the two-orbital system in CI and BS languages.16 He has advocated the use of the
corresponding orbital transformation (COT) due to Amos and Hall17 to determine the non-

18
Chapter 2 Theoretical Background

orthogonal, ‘valence-bond’-like magnetic orbital pairs.


It is evident that BS is not the only way to approach the problem of magnetic exchange
coupling at DFT level. Recently Filatov et al.9a-b proposed a methodology based on the
Restricted Ensemble Khon-Sham formalism which deals directly with spin eigenfunctions.
The procedure suffers from the strong dependence of exchange interaction with the exchange
correlation potential.
We find that there are two theoretical aspects involved here. First, the existing literature is
based on the interpretation of the N-electron spin in the two-electron spin picture, that is, from
an attempt to provide an interpretation of the calculated results for an N-electron system in
terms of the Heisenberg two-spin Hamiltonian given by Eq. (2.1). The second aspect deals
with the nature of the wave function. An unrestricted procedure basically relies on a single-
determinantal wave function as the ground state configuration. The so-called BS approach is
based on the single determinant. Therefore, any analysis of the results computed by the BS
method must be based on such wave functions. Complications would still arise when the
highest occupied molecular orbitals (HOMO) are degenerate. This work is on diradicals with
non-degenerate HOMOs.

2.2. Theoretical background


The Heisenberg spin Hamiltonian is an effective Hamiltonian. It is normally written
with a base line that equals the energy of the lowest spin state. For a diradical, the lowest spin
state is a singlet. Nevertheless, a diradical is generally based on a pair of non-bonding
orbitals which implies that the highest occupied molecular orbitals (HOMO) would be either
non-degenerate (the single determinant representation) or degenerate. From two electrons
occupying four spin orbitals corresponding to a pair of HOMOs, it is possible to build a set of
triply degenerate configurations (triplet) and three singlet configurations. Therefore, one
needs to identify the singlet that forms the base line.

2.2.1. Single determinant approach


The single determinants for the singlet (S), broken-symmetry (B) and triplet (T) states
in the unrestricted formalism are written as
ΨS1=||η1(r1)α(s1) η/1(r2)β(s2) ... ηN/2(rN−1)α(sN−1) η/N/2(rN)β(sN)||,
ΨS2=|| η~ 1(r1)α(s1) η~ /1(r2)β(s2) ... η~ N/2+1(rN−1)α(sN−1) η~ /N/2+1(rN)β(sN)||,

19
Chapter 2 Theoretical Background

ΨB1=||ζ1(r1)α(s1) ζ/1(r2)β(s2) ... ζN/2(rN−1)α(sN−1) ζ/N/2+1(rN)β(sN)||, (2.7)


ΨB2=||ζ/1(r1)α(s1) ζ1(r2)β(s2) ... ζ/N/2+1(rN−1)α(sN−1) ζN/2(rN)β(sN)||,
ΨT1=||ξ1(r1)α(s1) ξ/1(r2)β(s2) ... ξN/2(rN−1)α(sN−1) ξN/2+1(rN)α(sN)||,
ΨT2=||ξ/1(r1)α(s1) ξ1(r2)β(s2) ... ξN/2(rN−1)β(sN−1) ξN/2+1(rN)β(sN)||.
In this work, we do not explicitly consider the density functional treatment as was
done by Ginsberg, Noodleman and Davidson. Instead, we put forward a new spin
Hamiltonian that is valid for N electrons. It holds so long as the single determinant picture is
retained, the HOMOs are nondegenerate, and the energy ordering for the lowest six states,
three singlets and a triplet, remains intact. The spatial orbitals η’s are mutually orthogonal and
similarly η/’s are mutually orthogonal. The orbitals η’s are not necessarily orthogonal to η/’s.
These sets {η} and {η/} are strictly determined from the unrestricted calculation. Similarly,
ζ’s are mutually orthogonal, ζ/’s are also mutually orthogonal, but ζ’s need not be orthogonal
to ζ/’s, and the sets {ζ} and {ζ/} are obtained from the BS calculations. The same situation is
valid for the triplet orbitals belonging to the sets {ξ} and {ξ/}. The sets {η}, { η~ }, {ζ} and

{ξ} are in general somewhat different in the valence domain. Similarly, the sets {η/}, { η~ /},

{ζ/}and {ξ/} differ from each other in the valence sector.


The single determinant energy expectation values are

ES 1 = Ecore
S1
+ [h( N / 2)α , ( N / 2)α + h( N / 2) β , ( N / 2) β ]
N / 2 −1
+ ∑ [{J aα , ( N / 2)α − K aα , ( N / 2)α + J aβ , ( N / 2)α }
(2.8)
a =1

+ {J aβ , ( N / 2) β − K aβ , ( N / 2) β + J aα , ( N / 2) β }]
+ J ( N / 2)α , ( N / 2) β
where
N / 2 −1
S1
Ecore = ∑ (haα ,aα + haβ ,aβ )
a =1
(2.9)
1 N / 2−1 N / 2−1
+ ∑ ∑ ( J aα , bα − K aα , bα + J aα , bβ + J aβ , bβ − K aβ , bβ + J aβ , bα ).
2 a =1 b=1
We similarly write

20
Chapter 2 Theoretical Background

ES 2 = Ecore
S2
+ [h( N / 2+1)α , ( N / 2+1)α + h( N / 2+1) β , ( N / 2+1) β ]
N / 2 −1
+ ∑ [{Jaα , ( N / 2+1)α − K aα , ( N / 2+1)α + Jaβ , ( N / 2+1)α }
a =1 (2.10)
+ {Jaβ , ( N / 2+1) β − K aβ , ( N / 2+1) β + Jaα , ( N / 2+1) β }]
+ J( N / 2+1)α , ( N / 2+1) β ,
EB = EB1 = EB 2
= Ecore
B1
+ [h(′N / 2)α , ( N / 2)α + h(′N / 2+1) β , ( N / 2+1) β ]
N / 2 −1
+ ∑ [{J a′α , ( N / 2)α − K a′α , ( N / 2)α + J a′β , ( N / 2)α } (2.11)
a =1

+ {J a′ β , ( N / 2+1) β − K a′ β , ( N / 2+1) β + J a′α , ( N / 2+1) β }]


+ J (/N / 2)α , ( N / 2+1) β ,
ET = ET 1 = ET 2
= Ecore
T1
+ [h(′′N / 2)α , ( N / 2)α + h(′′N / 2) β , ( N / 2) β ]
N / 2−1
and + ∑ [{J a′′α , ( N / 2)α − K a′′α , ( N / 2)α + J a′′β , ( N / 2)α } (2.12)
a =1

+ {J a′′α , ( N / 2+1)α − K a′′α , ( N / 2+1)α + J a′′β , ( N / 2+1)α }]


+ ( J (//N / 2)α , ( N / 2+1)α − K (//N / 2)α , ( N / 2 +1)α )
where the tilde signs and the primes have been used to indicate that the integrals over the S1,
S2, B and T orbitals are in general different from each other.

2.2.2. Two-determinant configurations


A linear combination of the BS determinants produces an approximation to the Ms = 0
component of triplet state

ΨT3′ = 2−1/2(ΨB1+ΨB2) (2.13)

with MS = 0 and similarly an approximation to the third singlet state

ΨS3′ = 2−1/2 (ΨB1−ΨB2). (2.14)

The corresponding energy values are

ET 3 = EB − K /
/ (2.15)

and

21
Chapter 2 Theoretical Background

ES 3 = EB + K /
/ (2.16)

where the quantity K/ is written as

K / =< B1core | B 2core > K ( N / 2),( N / 2+1)′ ,


(2.17)
K ( N / 2),( N / 2+1)′ =< ς N / 2 ς ′N / 2+1 | ς ′N / 2+1 ς N / 2 > .

In a two-configuration self-consistent-field (TCSCF) process, the orbitals in ΨT3′ (and ΨS3′)

would undergo relaxation to some extent to form configuration ΨT3 (and ΨS3) such that ET3′

changes into ET3 that is equal to ET (and ES3′ changes to ES3). One would like to obtain an

approximately correct measure of this change, which is discussed in the following.

2.2.3. Orbital perturbation theory


The Fock operator for state B1 is written as
N/2 N / 2 +1
F B1 = h + ∑ ( Jˆ ς − Kˆ ς )+ ∑ ( Jˆ ς ′ β − Kˆ ς ′ β ) . (2.18)
aα aα a a
a =1 a =1
( ≠N / 2 )

From the energy expression (2.15), one can determine the Fock operators relevant to the state
T3 in unrestricted formalism. In doing so, we make use of the “frozen core” like assumption
that B1core B 2core remains unchanged. The spatial parts of the orbitals, ς and ς ′ , are
varied subject to the orthogonality constraints ς a ς b = δ ab and ς a′ ς b′ = δ ab . We
consider δ ET 3 = 0 for arbitrary variations of ς ’s and ς ′ ’s. For the valence orbitals, we
need the additional constraint ς N / 2 ς N′ / 2+1 = 0. These variations are then coupled together
by Lagrange’s undetermined multiplier technique, and the multiplier matrix can be easily
shown to be hermitean. By diagonalizing the multiplier matrix, one finds Hartree-Fock
equations for the spatial functions ς and ς′ . For the B1' component of T3, the Fock

operators are found as

FBT3
1
≈ F B1
/ (19a)

for all orbitals ςiα and ςi′β , except for the orbitals ςN / 2α and ς(′N / 2+1) β for which

22
Chapter 2 Theoretical Background

FB1 T 3 ≈ F B1 + F / ,
/

(19b)
F / = − < B1core | B 2core > ( Kˆ ς + Kˆ ς ′ α ).
N / 2β N / 2+1

B1
The approximately equal to sign (≈) indicates that the expression for F holds with the
possibility of minor changes in the spatial functions from ς and ς ′ to ς and ς′ . The
perturbation F/ changes the spatial functions that can be used to form the broken symmetry
determinant B1' (and B2' ) in ΨT3. The first order energy correction is non-vanishing only for
the HOMOs. Thus the perturbed orbital energy values can be written as

ες iα = εςB1iα + ε ς(2)iα + ...,


(2.20)
ες i′β = ε ςB1i′β + ες(2)i′β + ...
for i =1, ..., N/2 –1, and

ες = ε ςB1 + ες(1) + ες(2) + ... ,


N / 2α N / 2α N / 2α N / 2α
(2.21)
ες ′ = ε ςB1′ + ε ς(1)′ + ε ς(2)′ + ...
N / 2+1β N / 2+1β N / 2+1β N / 2+1β

where

ε ς(1) = ε ς(1)′ = − < B1core | B 2core > K (/N / 2),( N / 2+1) = − K . (2.22)
N / 2α ( N / 2+1) β
/

The second order corrections to the orbital energies are given by

|< ς iα | Kς | ς N / 2α >|2
( N / 2+1)α
/

ε ς(2)iα = |< B1core | B 2core >|2 ,


ε ςB1iα − ε ςB1 α
N /2
(2.23)
|< ς i′β | Kς | ς ′N / 2+1β >|2
N / 2β
/

ε ς(2)i′β = |< B1core | B 2core >|2


ε ςB1i′β − ε ςB1′ β
N / 2 +1

for i =1, …, N/2 –1, and for the orbitals ςN / 2α and ς′N / 2 +1 β ,

N / 2+1 |< ς N / 2α | Kς | ς jα >|2


( N / 2+1)α
/

ε ς(2)
N /2
α = |< B1core | B 2core >|2 ∑ ε ςB1 − ε ςB1jα
,
j =1 α
( ≠ N / 2) N /2
(2.24)
N /2 |< ς ′N / 2+1β | Kς | ς ′j β >|2
N / 2β
/

ε ς(2)′
N / 2+1β
= |< B1core | B 2core >|2 ∑ ε ςB1′ − ε ςB1′ β
.
j =1 β
N / 2+1 j

Equation (2.24) is only applicable to the non-degenerate HOMO case.

23
Chapter 2 Theoretical Background

2.2.4. SCF energy


The component B1' (of ΨT3) is an eigenfunction of the operator
N
H (0) = ∑i =1
FBT13 (i ).
/ (2.25a)
( electrons )

This is comparable to the so-called Hartree-Fock Hamiltonian in the usual cases. The
corresponding many-body perturbation is
H ′ = H full − H (0) (2.25b)

where Hfull is the full Hamiltonian in coordinate space.

The zeroth-order energy for the B1' component (in T3) can be written as the sum of the
perturbed orbital energies, that is,
N /2 N/2+1
EB(0)1′ = [ ∑ ε ςBα1 +
a
∑ εςB′1β ]a
a =1 a′=1
( ≠ N / 2)

+ [ε ς(1) + ε ς(1)′ ] (2.26)


N / 2α N / 2+1β
N / 2 −1
+[ ∑ (ες(2)aα + ες(2)a′ β ) + ες(2)N / 2α + ες(2)(′ N / 2+1)β ] + ...
a =1

The zeroth order sum in this expression equals EB1(0) while the first order sum equals –2K. A
large number of terms in the second-order sum cancel each other, thereby leaving a residual
sum of only two terms,
N / 2 −1
[ ∑ (ες(2)aα + ες(2)a′ β ) + ες(2)N / 2α + ες(2)(′ N / 2+1)β ] =
a =1

|< ς N / 2α | Kς | ς N / 2+1α >|2


( N / 2+1)α
/

| < B1core | B 2core >|2 [ (2.27)


ε ςB1 α − ε ςB1 α
N /2 N / 2 +1

|< ς ′N / 2+1β | Kς | ς ′N / 2 β >|2


N / 2β
].
/

+
ε ςB1′ β − ε ςB1′ β
N / 2 +1 N /2

The denominators involved are of opposite signs. Therefore, the residual sum is quite small
in magnitude.

24
Chapter 2 Theoretical Background

The first order correction is given by

EB(1)1′ = EB(1)1 + 2 K , (2.28)


(1)
where E B1 is the first-order correction for the original determinant B1. As the SCF energy

is determined by energy up to the first order in many-body perturbation, we get

|< ς N / 2α | Kς | ς N / 2+1α >|2


( N / 2+1)α
/

EB1′ = EB1 + |< B1core | B 2core >|2 [


ε ςB1 α − ε ςB1 α
N /2 N / 2 +1
(2.29)
|< ς N′ / 2+1β | Kς | ς N′ / 2 β >|2
N / 2β
] + ...
/

+
ε ςB1′ β − ε ςB1′ β
N / 2 +1 N /2

Comparing (2.15), we find that the energy of T3 is given by

|< ς N / 2α | Kς | ς N / 2+1α >|2


( N / 2+1)α
/

ET 3 = ET 3 + < B1core | B 2core >


2
[
ε ςB1 − ε ςB1
/

α α
N /2 N / 2 +1
(2.30)
|< ς ′N / 2+1β | Kς | ς ′N / 2 β >|2
N / 2β
] + ...
/

+
ε ςB1′ β − ε ςB1′ β
N / 2+1 N /2

Similarly, the S3 singlet energy can be written as

|< ς N / 2α | Kς | ς N / 2+1α >|2


( N / 2+1)α
/

ES 3 = ES 3 + < B1core | B 2core >


2
[
ε ςB1 − ε ςB1
/

α α
N /2 N / 2 +1
(2.31)
|< ς ′N / 2+1β | Kς | ς ′N / 2 β >| 2
N / 2β
] + ...
/

+
ε ςB1′ β − ε ςB1′ β
N / 2 +1 N /2

These formulae apply in the case of non-degenerate HOMOs. The situation that arises in the

degenerate HOMO case is not addressed theoretically in this thesis, although in Chapter 7 we

investigate a system that has the characteristic feature.

A linear combination of S3 and T3 yields a pair of degenerate BS functions ΨB1' and

ΨB2' with energy EB' = EB1' = EB2'. The energy EB' is not directly obtainable from a quantum
chemical computation as the functions B1' and B2' are not self-consistent. As Eqs. (2.28) and
(2.29) show, the formula

25
Chapter 2 Theoretical Background

ES 3 = ET 3 + 2 K (2.32)
is approximately valid.
The main purpose of the derivation given in this section is to decide the energy
ordering, as illustrated in Figure 2.1. It is reasonable to expect the same energy ordering from
a DFT calculation. The DFT calculations are generally carried out by Hartree-Fock-Kohn-
Sham equation of the form18

[ hˆ1 + VC ( r ) + VXC (r )]φi ( r ) = ε iφi ( r ) , (2.33)

that involves the exchange-correlation terms in VXC(r). The conventional Hartree-Fock


method can be viewed as a limiting case where the correlation contribution is completely
neglected.18 In turn, the DFT formalism can be viewed as a treatment where the effect of the
correlation contribution leads to a modification of the J and K integrals from their Hartree-
Fock counterparts plus some additional corrective terms. While the calculated total energy
can vary largely from UHF to UB3LYP because of the accommodation of the correlation
energy, the relative energy for different configurations generally changes by a lesser amount,
thereby leaving the energy ordering of the lowest-lying configurations intact in most cases.

S2 S2

S1 S1

S3/
S3 T1,T2
B1,B2 T3/
K/ T3
T3/ B1,B2
T1,T2 /
T3 K
S3/
S3
( K/ > 0) ( K/ < 0)
Figure 2.1. Schematic illustration of the energetics in the case of nondegenerate HOMO’s. The wave
functions S3 and T3 are multiconfigurational. The UHF (UB3LYP) singlet calculations
generally lead to S1.

2.3. Spin-Hamiltonian Treatment

26
Chapter 2 Theoretical Background

2.3.1. Base line


~
As S1 and S2 involve stronger repulsions ( J ( N / 2 )α , ( N / 2 ) β and J( N / 2+1)α , ( N / 2+1) β are greater

than J/(N/2), (N/2+1)/+ K (/ N/2), ( N/2+1) / ) and S2 has population in a higher orbital compared to S1, one

obtains the energy ordering ES2 > ES1 > ES3. It is assumed here that the difference in one-
electron energies hN/2, N/2 and hN/2+1, N/2+1 does not alter the energy ordering of S1 and S3. If
the energy ordering changes, the BS method would break down as it works on the assumption
that the BS wave function has energy midway between the energies of the singlet and the
triplet. Arbitrary breakdowns from the systematics have not been observed, except when the
Bencini-Ruiz formula (2.5) applies. But ΨS3 is a two-determinant configuration. In other
words, the single determinant picture gives a poor representation of the singlet ground state of
the diradical. In practice, one computes from UHF or related techniques like UB3LYP a
singlet energy that is quite high. Following the energy ordering in Figure 1, the base line for
the spin Hamiltonian is determined by ES 3 , the energy of the pure state Ψ S3 that is basically
0 0

formed from the configuration ΨS3.

2.3.2. Spin Hamiltonian


We write the N-electron spin-projected form for the full Hamiltonian of the diradical
as
0
H spin = Pˆ 0 H full P̂ 0 (2.34a)

where

Pˆ 0 = ∑ Pˆ SI0 + ∑ Pˆ TJ0 + ∑ Pˆ QK
0
+ ... , (2.34b)
I J K

0
the symbols S, T, Q, etc. indicating singlet, triplet, quintet, etc. states. The operator P̂SI is the

projector for the Ith singlet eigenstate Ψ SI of the full coordinate-space Hamiltonian Hfull, and
0

so on. The eigenstates can be built up from the configurations S1, S2, etc. in an explicitly
carried out many-body treatment. We note the energy eigenvalue orderings
0
ES3 < ES1
0
< ES2
0
< ... (2.35a)
etc., besides the equality
0
ET1 = ET2
0
= ET3
0
= ES3
0
− 2J . (2.35b)

27
Chapter 2 Theoretical Background

A little rearrangement gives rise to the expression


0
H spin = ( ES3
0
− JS 2 )Pˆ 0 + ∑ 0 ˆ0
( ESI0 − ES3 )PSI
I( ≠ 3)
+ ∑ ( ETJ0 − ES3
0
+ 2 J )Pˆ TJ0 (2.36)
J>3
+ ∑ ( EQK
0
− ES3
0
+ 6 J )Pˆ QK
0
+ ...
K
where S2 is the squared spin angular momentum operator. The effective spin Hamiltonian is
0
obtained by replacing the projector P̂ in Eq. (2.36) by the unit operator:
N N G G
0 ˆ0
H spin = E0′ − J ∑ ∑ / Si ⋅ S j + ∑ ( ESI0 − ES3 )PSI
i=1 j =1 I( ≠ 3)
+ ∑ ( ETJ0 − ES3
0
+ 2 J )Pˆ TJ0 (2.37a)
J>3
+ ∑ ( EQK
0
− ES3
0
+ 6 J )Pˆ QK
0
+ ...
K
G
where Si is the spin operator for the ith electron, the prime over the second sum indicates j ≠

i, and the base E'0 is given by


3N
E0′ = ES30 − J. (2.37b)
4
The Hamiltonian Hspin yields the eigenvalues for the respective eigenstates (that may be found
from an explicit many-body treatment using Hfull). We emphasize that N is the total number
of electrons involved in the calculation of energy and other molecular characteristics. The
retention of the same coupling constant J for all pairs of electrons is related to the
indistinguishability of the electrons.

Equation (2.37a) is to be distinguished from the Heisenberg spin Hamiltonian for


ferromagnetic and antiferromagnetic solids. The latter operator is written as
n n
G G
H FM / AFM = E0 − ∑∑ / J pq s p ⋅ sq
′′ (2.38)
p =1 q =1

28
Chapter 2 Theoretical Background

G G
where p and q are indices for the lattice sites, and s p and sq are the associated spins. Each

site has the same spin, say, s. Yamaguchi et al.15 used this Hamiltonian and considered s = ½
so that the maximum number of half spin is n with (n+1)-fold degeneracy. Thus the treatment
by Yamaguchi et al. has been an n-electron treatment where n is the number of monoradical
fragments.

For a diradical, there are only two sites, and in general no lattice is formed. For two sites,

Eq. (2.38) reduces to Eq. (2.1) and E0′′ equals E0. In the two-electron two-orbital model, Eq.

(2.37a) reduces to Eq. (2.1) where the need for the projectors are normally overlooked, A and
B represent two magnetically active electrons, and E0' becomes equal to E0. It is important to
note that by an appropriate partitioning of the core and valence spaces ascribed to each site,
the first two terms in Hspin of Eq. (2.37a) for any radical can be reduced to the ferromagnetic
and anti-ferromagnetic lattice Hamiltonian HFM/AFM in Eq. (2.38). Thus the first two terms in
Hspin together represent a general operator.

2.3.3. Expectation values


G N G
The total spin is written as S = ∑ S i such that
i =1

N N N G G
S 2 = ∑ S i2 + ∑ ∑ / S i ⋅ S j . (2.39)
i =1 i =1 j=1
G G
To evaluate the effect of the S i ⋅ S j terms in (2.34) and (2.37), one makes use of the equality
G G
Si ⋅ S j = (Si+Sj− + Si−Sj+)/2 + SizSjz. We define the overlap integrals as

kl =< η k | η l > ,
SS1 /

~ ~/
kl =< η k | η l > ,
SS2

S klB1 =< ζ k | ζ l/ > , (2.40)

kl =< ξ k | ξ l > ,
S T1 /

and find the following interesting expectation values of <S2>:


N N/ 2 N/ 2 S1 2
< S 2 > S1 = − ∑∑ | S kl | ,
2 k =1 l =1

29
Chapter 2 Theoretical Background

N/ 2+1 N/ 2+1
N
< S 2 >S2 =
2
− ∑ ∑| S
k =1 l =1
S2 2
kl| ,
( k ≠ N / 2) (l ≠ N / 2)

N 1 N/ 2 N/ 2 +1
1 N/ 2 +1 N/ 2
< S >S3 = − ∑
2
∑ |S | −
B′ 2
∑ ∑| S B′ 2
|
2 2 k =1 2
kl lk
l =1 k =1 l =1
( l ≠ N / 2) ( k ≠ N / 2)

− B1core | B 2core SBN′/ 2,( N / 2) SBN′/ 2+1,( N / 2+1) ,


/ /

N/ 2 +1
N N/ 2
< S 2 > B1 = −∑ ∑| S B 2
kl | , (2.41)
2 k =1 l =1
(l ≠ N/ 2 )

N/ 2 +1 N/ 2 -1
N
< S 2 > T1 = +1− ∑ ∑| S T 2
| ,
kl
2 k =1 l =1

N 1 N/ 2 N/ 2 +1
1 N/ 2 +1 N/ 2
< S >T3 = − ∑
2
∑ |S | −
B′ 2
∑ ∑| S B′ 2
|
2 2 k =1 2
kl lk
l =1 k =1 l =1
( l ≠ N / 2) ( k ≠ N / 2)

+ B1core | B 2core SBN′/ 2,( N / 2) SBN′/ 2+1,( N / 2+1) .


/ /

The condition SklS/B/T = δkl implies <S2>S1,2,3 = 0, <S2>B = 1 and <S2>T = 2. These ideal B

values are rarely obtained from unrestricted calculations on the BS and T states.
Strictly speaking, the S2 operator is not well-defined in DFT. The assumption that
STkl ≈ SBkl generally gives the difference <S2>T1 − <S2>B1 – 1 = |SBN/2,N/2+1| 2. The commonly
perceived difference 1−<S2>B1 = |SBN/2, N/2+1|
2
evolves from the very restrictive conditions

STkl = δ kl . These assumptions do not necessarily hold in every case. In fact, as Eqs. (2.41)
show, it is possible for the calculated <S2>B1 and <S2>T1 to be greater than 1 and 2
respectively, and their difference can be less than 1. Such results are often obtained from
calculations.

2.4. Coupling Constant


The SCF energy values can be obtained as the expectation values of the spin
Hamiltonian. These are generally written as

30
Chapter 2 Theoretical Background

⎛ 3N ⎞
E = E0′ + J ⎜ − < S2 > ⎟ + ∑ ( ESI0 − ES30 ) < Pˆ SI0 >
⎝ 4 ⎠ I( ≠ 3)

+ ∑ ( ETJ0 − ES30 + 2 J ) < Pˆ TJ0 > (2.42)


J>3
+ ∑ ( EQK
0
− ES30 + 6 J ) < Pˆ QK
0
> +...
K
An equivalent expression is

⎛ 3N ⎞
E = E0′ + J ⎜ − 2 < Pˆ T0 > ⎟ + ∑ ( ESI0 − ES30 ) < Pˆ SI0 >
⎝ 4 ⎠ I( ≠ 3)

+ ∑ ( ETJ0 − ES30 ) < Pˆ TJ0 > (2.43)


J>3
+ ∑ ( EQK
0
− ES30 ) < Pˆ QK
0
> +...
K
ˆ 0 = Pˆ 0 + Pˆ 0 + Pˆ 0 .
where PT T1 T2 T3

When the basis is large enough, the expectation values of S2 are nearly equal to the
ideal values 0, 1 and 2 for singlet, BS and triplet configurations. The addition of the sums on
the right side of (2.43) remains more or less the same for S3 and T3 (as the contribution from
the higher energy states is nearly equal in the two cases), and therefore for B1 and B2. One
also expects more or less the same sum for T1 and T2 on a similar ground. This happens
especially in the DFT calculations where the total of the sums is small. We write
3 NJ
E0′ +
4
+ ∑ ( ESI0 − ES30 ) < Pˆ SI0 > + ∑ ( ETJ0 − ES30 ) < Pˆ TJ0 >
I( ≠ 3) J>3 (2.44)
+ ∑ ( E 0 − E 0 ) < Pˆ 0 > +... = E
QK S3 QK c

K
where Ec remains practically same for the S3, T1 (T2), T3 and B1 (B2) configurations, but
changes with the basis set. The quantity Ec can be interpreted as ES3. In the limit of an
infinitely large basis where basis set truncation error is negligibly small, Ec approaches the
ES3 value in the DF (Hartree-Fock Kohn-Sham) limit.

It is possible to rewrite Eqs. (2.42) and (2.43) in terms of Ec as

31
Chapter 2 Theoretical Background

E = Ec − J < S 2 > +2 J ∑
< Pˆ TJ0 > +6 J ∑ < Pˆ QK
0
> +... , (2.45a)
J>3 K
and equivalently

E = Ec − 2 J < Pˆ T0 > (2.45b)

for a particular basis set. Equation (2.45a) shows that a fair estimate of J can be obtained as
( ELow spin − EHigh spin )
J= (2.46)
(< S 2 > High spin − < S 2 > Low spin )
that is the same as Eq. (6) derived by Yamaguchi from the energy expression for the two-
electron model of a diradical [and the n-electron model of a n-radical]. The high spin state is
defined here as the state that correspond to S = S1+S2. This formula, though approximate,
holds irrespective of the limit of the two-orbital overlap and irrespective of the number of
magnetically active orbitals, and remains valid as long as the single determinantal picture is
retained for B1 (B2) and T1 (T2), and the two-determinant configuration S3 (T3) is valid.

A better relation is obtained from Eq. (45b),


( ELow spin − EHigh spin )
J= , (2.47)
2(< Pˆ 0 >
T
− < Pˆ 0 >
High spin T Low spin
)
but it will be difficult to use this relation in the absence of the knowledge of the exact states
Ψ 0T1 , Ψ T2
0
and Ψ T3 . This relation is in reality a spin projection formula. When the
0

ˆ > and < Pˆ > equals 1/2, one obtains the GND equation, (2.4).
difference between < P
0 0
T T1 T B1

If one uses the condition SklS/B/T = δkl that is reminiscent of the restricted scheme, one
obtains from Eq. (2.45b), ES3 = Ec, EB1 ≈ Ec−J, and ET1 ≈ Ec−2J such that ES3− ET1 ≈ 2J and
EB1 − ET1 ≈ J. Considering SN/2, (N/2+1) ≡ Sab to be a small but finite quantity, one immediately
gets the Noodleman equation (2.2). Actually, the Noodleman equation requires only one
assumption, that is, while SBN/2, N/2+1 may or may not equal zero, the sum of the squares of the
rest of the overlap integrals for B1 in (2.41) is more or less equal to the sum of the overlap
integral squares for T1 in the same equations. It does not require the restricted Hartree-Fock
type constraints.
The deviation of J from (2.46) mainly occurs due to the spin correlation effects as
32
Chapter 2 Theoretical Background

shown in Eq. (2.45a). For a smaller basis set, the quantity Ec may vary for different
configurations. In such a case, both (2.46) and (2.47) would give rise to a large fractional
deviation in the calculated value of J. Most of the organic diradicals have a small coupling
constant, and the deviation calculated by using Yamaguchi or GND expressions and a smaller
basis set may not be noticeably large, though it is there. As Eq. (2.44) shows, the lack of
constancy of Ec is a general correlation effect.

2.5. Discussion
A spin Hamiltonian can also be derived from effective Hamiltonian theory of Illas et
al. 7 Here, the spin Hamiltonian is obtained for N electrons.

2.5.1. Factors Influencing Accuracy


A number of factors affect the accuracy of a calculation of the magnetic exchange
coupling constant within the BS approach. First of all, the DF methodology is to be
employed. A simple UHF calculation may yield different Ec values for S3, T3, B1 (B2), and
T1 (T2). Second, the basis must be large in size and must contain polarization and diffuse
functions so as to reduce the basis set truncation error. Third, the computed values of <S2>B1
and <S2>T1 need to be as close to 1.00 and 2.00 respectively as possible. A large deviation of
the S2 expectation value indicates the build-up of inaccuracy in the computed energy values,
and then the use of Eqs. (2.46) and (2.47) becomes suspect. Fourth, correction terms can be
added to the spin Hamiltonian defined in (2.37a). Spin biquadratic correction terms have
been considered by Noodleman et al.3c However, as these terms arise only in an indirect way
and not from a direct spin dipole−spin dipole interaction, their contribution to J is expected to
be rather small. Higher order terms in the spin Hamiltonian have been important both
experimentally as theoretically for superconducting couplers.7g

2.5.2. Numerical Tests


The S2-weighted projection

W = 2∑ < Pˆ TJ0 > +6∑ < Pˆ QK


0
> + ...
J >3 K >3 (2.48)
= < S > −2 < Pˆ >
2 0
T

33
Chapter 2 Theoretical Background

is a crucial quantity. The J value calculated from Eq. (2.46) differs from that calculated by
Eq. (2.47) by a factor of 1/[1−(WB1 – WT1)]. A large ΔW indicates that the Yamaguchi
expression is in error. When the basis is sufficiently large, a measure of W can be obtained
ˆ > B1 ≈0.5 and < Pˆ > T1≈1.0 that leads to Eq. (2.4). These
from the approximations < P
0 0
T T

yield
Ec (B1, T1) = 2 EB1 − ET1 (2.49)

that can be visualized as the average value of ES3 estimable from the computed total energies.
We consider two typical nitronyl nitroxide diradicals, namely, (i) D-NIT2 and (ii) 2,2′-
(1,2-ethynediylid-4,1 phenylene) bis [4,4,5,5–tetramethyl–4,5-dihydro-1H-imidozolyl-oxyl].
These are shown in Figure 2.2. The basic data have been taken from our previous work19,20
using the UB3LYP methodology on Gaussian98 software.21 The quantities W, J and Ec are
calculated here. The results are given in Tables 2.1 and 2.2 respectively. The observed J
values are 349.6 cm−1 for D-NIT2 in solid22 and –3.37 cm−1 for species (ii) in solution.23 The
calculations presented in this section are only to illustrate the performance of the present
methodology rather than to provide accurate values.

Table 2.1. Spin-weighted projection, coupling constant and estimated energy of the singlet
configuration from single-point UB3LYP calculations on D-NIT2.a

J in cm−1 c
b
W Ec in a.u.b
Basis B1 T1 Eq. (46) Eq. (47) b Eq. (49)

6−31+G** 0.1273 0.0620 −375.1 −350.6 −1145.0839


**
6−311+G 0.1286 0.0629 −375.2 −350.5 −1145.3303
6−311++G** 0.1285 0.0629 −375.0 −350.4 −1145.3305

a
Geometry optimization at ROHF/6-311G** level, ref. 19.
b
Assuming PˆT = 0.5, PˆT = 1.0.
B1 B1
c
Jobs=350.2 cm−1, ref . 22.

For species (i), the molecular geometry was optimized by ROHF method using 6-
311G** basis set. Both singlet and triplet optimized geometries are close to each other
34
Chapter 2 Theoretical Background

and also close to the crystal geometry. The WB1 and WT1 values remain more or less
unchanged while the basis set is changed, but they differ from each other (Table 2.1).
Consequently, the value calculated by the Yamaguchi formula is almost 7% larger than the J
value calculated from (2.47). The bases here are large enough to yield a near constancy of the
calculated J values as the basis size increases. Equation (2.47) obviously yields almost the
exact experimental J. The approach of Ec to a limiting value is manifest in this table.
The second molecule is quite large, and its molecular geometry was optimized at the
UHF level using the 6-31G(d) basis set. Single-point calculations were performed with
higher basis sets. In this case we find almost the same values for WB1 and WT1 such that ΔW≈
0 (Table 2.2). Thus we have more or less the same J value calculated from Eq. (2.46) and Eq.
(2.47). This fortuitous result presumably arises from the linear geometry enforced by the
acetylenic bond and the p-phenylene couplers. The coupling constant is antiferromagnetic
and very small in magnitude. The fractional variation of J with the basis size is significantly
large, and it is obvious that a large basis is required to calculate a reasonably accurate J value.
The largest basis set used is 6-311G(d,p), and the correspondingly calculated J value becomes

(i)

(ii)
Figure 2.2. The diradicals (i) D-NIT2 and (ii) 2,2′-(1,2-ethynediylid-4,1 -phenylene) bis
[4,4,5,5 –tetramethyl -4,5-dihydro-1 H-imidozolyl-oxyl].

35
Chapter 2 Theoretical Background

somewhat larger than the observed one. This has happened because the geometry was
optimized at a lower level. In any case, the approach of Ec towards a limiting value is
apparent though the limiting value itself is not manifest.

2.6. Conclusions
An analysis is made of the single determinant approach for a diradical in the non-
degenerate HOMO case. In general, one computes the determinants S1, B1 (B2) and T1 (T2).
Another determinant S2 is in existence. The determinants B1 and B2 can linearly combine to
produce the configurations S3/ and T3/. A two-configuration SCF will lead to S3 and T3 that
are slightly perturbed version of S3/ and T3/. The total energy for S3 (T3) differs from that
for S3/ (T3/) by a second order correction. The energy ordering is expected to S3 < S1 < S2.

Table 2.2. Spin-weighted projections, coupling constant and estimated energy of the singlet
configuration from single-point UB3LYP calculations on Species (ii).a
b
W J in cm−1 c Ec in a.u. b
Basis B1 T1 Eq. (46) Eq. (47) b Eq. (49)

6−31G** 0.0228 0.0225 −1.52 −1.51 −1455.5768


6−31+G** 0.0247 0.0241 −1.87 −1.87 −1455.6214
*
6−311G 0.0229 0.0221 −2.85 −2.86 −1455.8490
6−311G** 0.0226 0.0227 −3.60 −3.60 −1455.8949

a
Geometry optimization at UHF/6-31** level, ref. 20.
b
Assuming PˆT = 0.5, PˆT = 1.0.
B1 B1
c −1
Jobs= −3.37 cm , ref. 23.

An N-electron effective spin Hamiltonian is formulated. The base line for the spin

Hamiltonian is determined by ES3 that is the energy of the state Ψ S3 , an eigenstate of the full
0 0

Hamiltonian in coordinate space. The relationship with the Heisenberg two-spin Hamiltonian
and the Yamaguchi Hamiltonian can be established. An expression for the energy expectation
values is easily obtained. From this, we derive two expressions (2.46) and (2.47) for J, when
36
Chapter 2 Theoretical Background

the DFT methodology is adopted and the basis set is large. In general, Eq. (2.46) is
approximate in nature. It is identical with the expression due to Yamaguchi. Equation (2.47)
is more correct. When the basis is very large, (2.47) reduces to the so-called GND formula.
An average estimate of ES3 is also obtained. A smaller basis produces a large fractional
deviation for J.
These expressions are investigated by considering two nitronyl nitroxide diradicals. In
one case, the Yamaguchi approximation differs from the spin projection formula, both
yielding almost unvarying J values for different bases, and a limiting value of ES3 is observed.
In the other case, the Yamaguchi and spin projection give the same J, but the J value
improves with an increasing basis size. For both the species, the spin projection formula leads
to the observed value of J.

2.7. References
1 (a) Nesbet, R. K. Ann. Phys. 1958, 4, 87. (b) Nesbet, R. K. Phys. Rev, 1960, 119, 658. (c) Anderson, P.
W. Phys. Rev. 1959, 115, 5745. (d) Anderson, P. W. Solid Sate Phys. 1963, 14, 99. (e) Herring, C.
Magnetism, Rado, G. T., Shul, H., Eds.; Academic Press : New York, 1965; Vol. 2B. (f) Maynau, D.;
Durand, Ph.; Daudey, J. P.; Malrieu, J. P. Phys. Rev. A 1983, 28, 3193.
2 (a) de Loth, Ph.; Cassoux, P.; Daudey, J. P.; Malrieu, J. P. J. Am. Chem. Soc. 1981, 103, 4007. (b)
Miralles, J.; Castell, O.; Caballol, R. Chem. Phys. 1994, 179, 377. (c) Wang, C.; Fink, K,; Staemmler,
V. Chem. Phys. 1995, 192, 25. (d) Calzado, C. J.; Cabrero, J.; Malrieu, J. P.; Caballol, R. J. Chem.
Phys. 2002, 116, 3985. (e) Calzado, C. J.; Cabrero, J.; Malrieu, J. P.; Caballol, R. J. Chem. Phys. 2002,
116, 2728. (f) Ciofini, I.; Daul, C. A. Coord. Chem. Rev. 2003, 238-239, 187.
3 (a) Noodleman, L. J. Chem. Phys. 1981, 74, 5737. (b) Noodleman, L.; Baerends, E. J. J. Am. Chem.
Soc. 1984, 106, 2316. (c) Noodleman, L.; Davidson, E. R. Chem. Phys. 1986, 109, 131. (d) Noodleman,
L.; Peng, C. Y.; Case, D. A.; Mouesca, J.-M. Coord. Chem. Rev. 1995, 144, 199.
4 Ruiz, E.; Alemany, P.; Alvarez, S.; Cano. J.; J. Am. Chem. Soc. 1997, 119, 1297.
5 Bencini, A.; Gatteschi, D.; Totti, F.; Sanz, D.N.; Mc Clevrty, J. A.; Ward, M. D. J. Phys. Chem. A 1998,
102, 10545.
6 Ross, P. K.; Solomon, E. I. J. Am. Chem. Soc. 1991, 113, 3246.
7 (a) Martin, R. L.; Illas, F. Phys. Rev. Lett. 1997, 79, 1539. (b) Caballol, R.; Castell, O.; Illas, F.;
Moreira, I. di P. R.; Malrieu, J. P. J. Phys. Chem. A 1997, 101, 7860. (c) Barone, V.; Matteo, A. di;
Mele, F.; Moreira, I. di P. R.; Illas, F. Chem. Phys. Lett. 1999, 302, 240. (d) Illas, F.; Moreira, I. di P.
R.; Graaf, C. De; Barone, V. Theor. Chem. Acc. 2000, 104, 265. (e) Graaf, C. de; Sousa, C.; Moreira, I.
di P. R.; Illas, F. J. Phys. Chem. A 2001, 105, 11371. (f) Moreira, I. de P. R.; Calzado, C. J.; Malrieu ,

37
Chapter 2 Theoretical Background

J. P.; Illas, F. Phys. Rev. Lett, 2006, 97, 087003. (g) Moreira, I. de P.R.; Suaud, N.; Guihéry, N.;
Malrieu, J.P.; Caballol, R.; Bofill, J.M.; Illas, F. Phys. Rev. B 2002, 66, 134430.
8 Rodriguez-Fortea, A.; Alemany, P.; Alvarez, S.; Ruiz, E. Inorg. Chem. 2002, 41, 3769.
9 (a) Illas, F.; Moreira, I. di P. R.; Bofill, J. M.; Filatov, M. Phys. Rev. B 2004, 70, 132414. (b) Illas, F.
Moreira, I. de P. R.; Bofill, J. M.; Filatov, M. Theoret. Chem. Acc., 2006, 115, 587. (c) Dai, D.;
Whangbo, M-H. J. Chem. Phys. 2003, 118, 29.
10 (a) Ruiz, E.; Alemany, P.; Alvarez, S.; Cano. J. Inorg. Chem. 1997, 36, 3683. (b) Ruiz, E.; Cano. J.;
Alvarez, S.; Alemany, P. J. Am. Chem. Soc. 120, 11122. (c) Ruiz, E.; Alvarez, S.; Alemany, P. Chem.
Com. 1998, 2762. (d) Ruiz, E.; Cano. J.; Alvarez, S.; Alemany, P. J. Comput. Chem. 1999, 20, 1391. (e)
Yamaguchi, K.; Takahara, Y.; Fueno, T.; Nasu, K. Jpn. J. Appl. Phys. 1987, 26, L1362. (f) Onishi, T.;
Soda, T.; Kitagawa, Y.; Takano, Y.; Daisuke, Y.; Takamizawa, S.; Yoshioka, Y. Yamaguchi, K. Mol.
Cryst. Liq. Cryst. 2000, 143, 133.
11 Ginsberg, A. P. J. Am. Chem. Soc. 1980, 102, 111.
12 Noodleman, L.; Davidson, E. R. Chem. Phys. 1986, 109, 131.
13 Ruiz, E.; Cano. J.; Alvarez, S.; Alemany, P. J. Comput. Chem. 1999, 20, 1391.
14 Bencini, A.; Totti, F.; Daul, C. A.; Doclo, K.; Fantucci, P.; Barone, V. Inorg. Chem. 1997, 36, 5022.
15 (a) Yamaguchi, K.; Fukui, H.; Fueno, T. Chem. Lett. 1986, 625. (b) Yamaguchi, K.; Takahara, Y.;
Fueno, T.; Nasu, K. Jpn. J. Appl. Phys. 1987, 26, L1362. (c) Yamaguchi, K.; Jensen, F.; Dorigo, A.;
Houk, K. N. Chem. Phys. Lett. 1988, 149, 537. (d) Yamaguchi, K.; Takahara, Y.; Fueno, T.; Houk, K.
N. Theo. Chim. Acta 1988, 73, 337.
16 Neese, F. J. Phys. Chem. Solid 2004, 65, 781.
17 Amos, A.T. Hall, G.G. Proc. R. Soc. Ser. A. 1961, 263, 483.
18 Ruiz, E.; Alvarez, S.; Cano, J.; Polo, V. Chem. Phys. 2005, 123, 164110.
19 Vyas, S.; Ali, Md. E.; Hossain, E.; Patwardhan, S.; Datta, S. N. J. Phys. Chem. A 2005, 109, 4213.
20 Ali, Md. E.; Vyas, S.; Datta, S. N. J. Phys. Chem. A 2005, 109, 6272.
21 Gaussian 03, Revision C.02, M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J.
R. Cheeseman, J. A. Montgomery, Jr., T. Vreven, K. N. Kudin, J. C. Burant, J. M. Millam, S. S.
Iyengar, J. Tomasi, V. Barone, B. Mennucci, M. Cossi, G. Scalmani, N. Rega, G. A. Petersson, H.
Nakatsuji, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda,
O. Kitao, H. Nakai, M. Klene, X. Li, J. E. Knox, H. P. Hratchian, J. B. Cross, C. Adamo, J. Jaramillo,
R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, P. Y.
Ayala, K. Morokuma, G. A. Voth, P. Salvador, J. J. Dannenberg, V. G. Zakrzewski, S. Dapprich, A. D.
Daniels, M. C. Strain, O. Farkas, D. K. Malick, A. D. Rabuck, K. Raghavachari, J. B. Foresman, J. V.
Ortiz, Q. Cui, A. G. Baboul, S. Clifford, J. Cioslowski, B. B. Stefanov, G. Liu, A. Liashenko, P.
Piskorz, I. Komaromi, R. L. Martin, D. J. Fox, T. Keith, M. A. Al-Laham, C. Y. Peng, A.

38
Chapter 2 Theoretical Background

Nanayakkara, M. Challacombe, P. M. W. Gill, B. Johnson, W. Chen, M. W. Wong, C. Gonzalez, and J.


A. Pople, Gaussian, Inc., Wallingford CT, 2004.
22 Ziessel, R.; Stroh, C.; Heise, H.; Köhler, F. K.; Turek, P.; Claiser, N.; Souhassou, M. ; Lecomte, C.
J. Am. Chem. Soc. 2004, 126, 12604.
23 Wautelet, P.; Moigne, J. Le ; Videva, V.; Turek, P.; J. Org. Chem. 2003, 68, 8025.

39
Chapter 3
Organic Fused─Ring Diradicals

This Chapter describes the investigation of ground state spins of seven diradicals belonging to the
fused ring system by ab initio restricted and unrestricted formalisms. In this work a variety of basis sets
is used. The UHF calculations yield an unrealistically large Singlet−Triplet (S−T) splitting. To avoid
spin contamination completely, we have repeated computations in the restricted (open-shell) Hartree-Fock
framework. The R(O)B3LYP/6-311G(d,p) optimized geometry yields the best total energy for each spin
state and hence the most reliable S−T energy difference. The calculated results are in agreement with the
available experimental findings. Molecules 3 and 7 have widely different geometries in the singlet and
triplet states. The UHF spin density plots obtained from the 4-31G optimized geometries manifest the
phenomenon of spin alternation in the ground state.
Chapter 3 Organic Diradicals

3.1. Introduction
Non-Kekulé hydrocarbons are known to be diradicals and highly reactive.1 The
presence of degenerate nonbonding molecular orbitals (NBMOs) is responsible for their
extraordinary reactivity.2 A singlet ground state results when the degeneracy is spoiled. In
fact, Hoffmann has shown that when the NBMOs differ by less than 1.5 eV, the ground state
is a triplet.3 It is also well-known that a change of the symmetry of the molecule or a
variation of the electronegativity of the diradical termini can be used to control the spin
multiplicity in the ground state.4 These multiplicities can be reliably predicted by ab-initio
post-Hartree-Fock treatments using large basis sets.5
In a previous work Datta et. al.6 discussed the spin nature of some chain and
monocyclic diradicals, and found the UCCSD(T) methodology with split-valence basis sets to
be a dependable approach to the calculation of the S−T energy difference for diradicals. In
this work we use ab-initio methods to characterize the ground state spin multiplicity of seven
diradicals sharing the naphthalene skeleton (1-7). Among these molecules, one is a 1,3-
substituted naphthalene (1) and the rest are 1,8-substituted naphthalene derivatives (2-7).
These molecules are shown in Figure 1.1. Out of these, 2 and 5 are derivatives with
homonuclear substituents, and 1, 3, 4, 6 and 7 have heteroatom substituents. These non-
Kekulé diradicals exhibit very high reactivity.4 The species 4-oxy-2- naphthalenyl methyl (1)
was observed by ESR.7 While plausible zwitterionic singlet structures may be drawn for
species 1, a Curie law analysis showed that it has a triplet ground state.4 Molecule 1,8-
naphthalenediylbis(methyl) (2) was observed by Pagni et al.8 using the triplet ESR spectrum.
The molecule was postulated to be a ground state triplet.4 INDO calculations for the planar,
anti-conformation of 8-imino-1-naphthalenyl methyl (3) have been carried out by Platz et al.4
Molecule 1,8- naphthalenediylbis(amidogen) (4) was observed by Platz et al.4 from ESR and
subsequent Curie law plot. The species 8-methyl-1-naphthyl carbene (5) was prepared by the
photochemical reaction of 8-methyl-1-naphthyl diazomethane at 4K. Platz et al.4 predicted
molecule 5 to be a ground state triplet based on the observation by Trozzolo et al.9 that 1-
naphthyl carbene is a ground state triplet. The species 8-methyl-1-naphthalenyl imidogen (6)
was matrix isolated by Platz and Burns.10
An accurate calculation of the multiplet splittings in non-Kekulé systems is a
challenging task. Based on the calculations on trimethylene methane (TMM), Borden,
Davidson and Feller11 had initially shown that the restricted (open-shell) Hartree-Fock
40
Chapter 3 Organic Diradicals

Figure 3.1. Species investigated in this work: (1) 4-oxy-2-naphthalenyl methyl, (2) 1,8-
naphthalenediylbis(methyl) or 1,8-naphthaquinodimethane (1,8-NQM), (3) 8-imino-1-
naphthalenyl methyl, (4) 1,8- naphthalenediylbis(amidogen), (5) 8-methyl-1-naphthyl
carbene, (6) 8-methyl-1-naphthalenyl imidogen and (7) 8-methyl-1-naphthyl
diazomethane.

[R(O)HF] and two-configuration self-consistent-field (TCSCF) calculations generally fail to


produce the correct relative energies and geometries although they may provide qualitatively
correct molecular orbitals for the two open−shell electrons in a diradical. This is a result of
the so-called doublet instability problem in RHF which is most severe when the basis set is
small.
Hence, these authors advocated the use of the unrestricted Hartree-Fock (UHF)
methods for a reasonably correct description of triplet and open-shell singlet geometries. In
this area, the most detailed investigations were carried out by Cramer and Smith,12a Nachtigall
and Jordan,13c and Mitani et al.14h on the molecules TMM, tetramethylene ethane (TME) and
m-xylylene respectively. They all concluded that the Singlet−Triplet energy differences are
highly sensitive to the methodology and the rigor (basis set) employed in the calculations.
41
Chapter 3 Organic Diradicals

The most widely studied diradical systems are TMM,12a-k TME,13a-j m-xylylene,14 and
polycyclic π-conjugated hydrocarbon polymers.15
Species 1-4 and 6 in Figure 1 were shown (and 5 was predicted) to be triplet in the
ground state. 4, 7-10 Species 7 is an exception. The traditional view of the chemist would be to
put one lone electron on the –CH2 substituent and the other on the two nitrogen
atoms of the substituent –CH2N2 in 7. This would indicate, by the rule of spin alternation,
that molecule 7 should be a ground state singlet, or at best a ground state triplet with very
little Singlet-Triplet energy gap. The main objective of this work is to confirm these
observations and generate quantitative data for the S−T energy gap.

3.2. Methodology
In this work we deal with diradical systems that are fairly large in size, and a complete
geometry optimization is required for each species in each spin state at both UHF and
R(O)HF levels. The STO-3G, 4-31G, 6-311G(d) and 6-311G(d,p) basis sets have been
employed in these calculations using the software Gaussian-98, but only the 6-311G(d) and 6-
311G(d,p) results are explicitly shown here. Density functional (DFT) calculations have been
performed by both UB3LYP and R(O)B3LYP methods while using some of the above-
mentioned basis sets. For the post-Hartree-Fock calculations at the unrestricted coupled-
cluster UCCSD and UCCSD(T) levels as well as the unrestricted density functional treatment
at UB3LYP level, the UHF-optimized geometry has been used for each species in each spin
state. The coupled cluster calculations get the sign right and never overestimate the gap.
Sometimes the gap is seriously underestimated. The spin-squared expectation values (which
should be 2.0) fall in a narrow range 3.2 to 3.5. Therefore, the coupled-cluster results are not
shown here. A complete geometry optimization was carried out at the DFT level only in the
restricted formalism using the R(O)B3LYP method.
A correct description of low-lying singlet and triplet states requires the proper
treatment of both static and dynamic correlation energy. The choice of UCCSD and
UCCSD(T) post-Hartree-Fock treatments for the present set of calculations is quite reasonable
from this standpoint. Using m-xylylene, Mitani et al.14h showed that the triplet state tends to
overstabilise relative to the singlet in a simple UHF calculation, whereas Møller-Plesset (MP)
perturbation calculations result in the singlet state being much more stable than the triplet.
Due to the near degeneracy of more than one UHF wave function, the MP perturbation theory

42
Chapter 3 Organic Diradicals

fails to yield correct results for diradical species. The zeroth-order UHF Hamiltonian is not
properly represented in such cases, thereby decreasing the credibility of the ensuing
perturbation-theoretic expansion. Coupled-cluster (CC),14h multiconfiguration self-consistent-
field (MCSCF)12a,13c and complete active space perturbation theory (CASPT2N)12a
methodologies treat correlation in a more sensible way, and by employing these calculational
procedures with progressively higher level of sophistication, one can overcome the problem
of one state being relatively more stabilized than the other. The importance of these
methodologies can be visualized from Table 2 in ref. 12a (for MCSCF and CASPT2N), Table
3 in ref. 13c (for MCSCF) and Table 3 in ref. 14h (for CC). Using TMM, Cramer and
Smith12a also demonstrated the restrictive nature of the density functional treatment (DFT).

The unrestricted methods like UHF, UCCSD, UCCSD(T), etc. introduce some bias
due to spin contaminations. Large deviations in the <S2> value raise questions on the
reliability of predictions made using the highly spin-contaminated geometries. There are
special methods like multi-reference coupled-cluster or non-standard version of DFT, which
are able to treat such systems reliably, concerning the interplay of static and dynamic
correlation effects.

The reason to perform computations in the restricted formalism is to correct the effect
of the spin-contaminated geometries in predicting the S−T gaps. Geometry optimization was
carried out at the STO-3G, 6-311G(d) and 6-311G(d,p) level using the R(O)HF methodology
and at the 6-311G(d,p) level using the R(O)B3LYP methodology.

3.3. Energy Differences


Ab-initio calculations on species 1-7 were performed at various levels, namely, UHF,
R(O)HF, UB3LYP, R(O)B3LYP, UCCSD and UCCSD(T), by employing STO-3G, split-
valence as well as a few polarized basis sets. In every case, the Singlet−Triplet energy gap
varies with the rigor of calculation as well as the basis set, and this observation is in
agreement with the trends noticed earlier in refs.7, 8, 9 and 15. The UHF methodology
generally yields the spin-contaminated geometry, and fails to give the correct Singlet−Triplet
energy gap. As the level of calculation increases, this gap generally converges. For reasons
discussed in the previous section, results at the MP level and the UCCSD(T) calculations have
not been shown here.
43
Chapter 3 Organic Diradicals

Cramer and Smith have demonstrated that the DFT methodology cannot adequately
account for the static correlation effects in closed-shell singlets in the limit of degenerate
frontier molecular orbitals, where one would expect multi-configurational behavior. The DFT
breakdown is expected for molecules 2, 4 and 6.

3.3.1. 4-oxy-2-naphthalenyl methyl


The molecule 4-oxy-2-naphthalenyl methyl (1) is planar with Cs symmetry in both
singlet and triplet states. Table 3.1 shows the energy values computed by the restricted and
the unrestricted formalisms, the <S2> value in the triplet state, and the point group. In all
calculations on this species, except the UB3LYP/6-311G(d) one, the triplet is found to be the
ground state. The best energy gaps obtained by our calculations, 9.7 kcal mol−1 at the
R(O)B3LYP/6-311G(d,p) level are in strong agreement with the CASPT2N/6-31G* energy
gap (11.6 kcal mol−1) reported by Hrovat et al.16

Table 3.1. The ab-initio total energy and the optimized geometry for the spin states of 4-oxy-2-
naphthalenyl methyl (1) in the unrestricted Hartree-Fock formalism. S and T indicate
singlet and triplet respectively.

ES− ET
Method Basis sets Optimization ET(a.u.) <S2>T
(kcal mol−1)

UHF 6-311G(d) O -496.1265 3.1744 39.1


UB3LYP 6-311G(d) SP -499.2106 2.0414 –14.7

ROHF 6-311G(d) O −496.0855 2.0000 13.3


ROHF 6-311G(d,p) O −496.1270 2.0000 30.3
ROB3LYP 6-311G(d,p) O −499.2665 2.0000 9.7

CASSCF 6-31G(d) O −496.1703 2.0000 14.2a


CASPT2N 6-31G(d) O −497.6266 2.0000 11.6a

a
Ref. 16.

The R(O)B3LYP calculations using a larger basis set including polarization functions
yields a more realistic S−T energy gap. Goodman and Kahn17 estimated, by using
photoacoustic calorimetry, that the energy difference is about 18.5 kcal mol−1. However, the
calculations by Hrovat et al.16 and those presented here indicate that this number is likely to
be an overestimation. For the other molecules too, the computed results have been given in

44
Chapter 3 Organic Diradicals

the same fashion. The UHF spin density is plotted using Hyperchem.18 All other calculations
are performed using Gaussian 98 software.19

Table 3.2. The ab-initio total energy and the optimized geometry for the spin states of
1,8−naphthalenediylbis(methyl) (2).

ES− ET
Method Basis sets Optimization ET(a.u.) <S2>T
(kcal mol−1)

UHF 6-311G(d) O –460.2974 3.1408 37.8


UB3LYP 6-311G(d) SP –463.3069 2.0938 16.1

ROHF 6-311G(d) O –460.2328 2.0000 37.8


ROHF 6-311G(d,p) O –460.2516 2.0000 37.8
ROB3LYP 6-311G(d,p) O –463.3178 2.0000 9.4

3.3.2. 1,8−naphthalenediylbis(methyl)
The molecule 1,8−naphthalenediylbis(methyl) (2) has C2v symmetry in both singlet
and triplet states. In all the calculations performed here the triplet has emerged as the ground
electronic state (Table 3.2), which agrees with the observation of Platz et al.4 and Pagni et al.8
The molecule seems to be a prime candidate for the DFT breakdown, but its NBMOs take
part in π−orbital formation, thereby making the system simultaneously planar and stable. The
R(O)B3LYP method yields the best result for the Singlet−Triplet splitting (9.4 kcal mol−1).

3.3.3. 8−imino−1−naphthalenyl methyl


The species 8−imino−1−naphthalenyl methyl (3) is a planar molecule with Cs point
group in both singlet and triplet states. Table 3.3 shows the computed Singlet−Triplet energy
gap by different methodologies. The molecule is found to be a ground state triplet in
accordance with the observation of Platz et al.4 who relied on INDO calculations. The INDO
method, however, is grossly inadequate to give rise to the correct S−T splitting. The splitting
calculated by Platz et al.4 was 60 kcal mol−1. The S−T energy gap follow the trends
mentioned earlier and the best value calculated in the present work is 16.4 kcal mol−1
[R(O)B3LYP/6-311G(d,p)].
The two substituents on the naphthalene ring (−CH2 and −NH) are of entirely different
nature. The methylene NBMO is mixed with the π orbitals. The −NH group has the lone
45
Chapter 3 Organic Diradicals

electron in a σ orbital. The most remarkable point about the optimized geometries is that the
singlet has −NH2 and −CH substituents whereas the triplet has −NH and −CH2 substituents, as
shown in Figure 3.2.
21 18 21 18
H 17
H H 16 H 20 H
H 17 N
N 20 16

15 3
14 H 1
H 1 H 15 H 14
3 2 19 13
2 19 13

4 4 8 11
8 11 6 9
9 5
6 H
H H 12 H 12
5
7
H H 10
H H 10
7

S T
Figure 3.2. Optimized singlet (S) and triplet (T) structure for 3.

Observing the optimized geometry, one may notice that the nitrogen is in the
state of sp2 hybridization in singlet and sp3 hybridization in triplet whereas the
substituent carbon is in sp3 hybridization in singlet and sp2 hybridization in triplet.
Thus, the optimized singlet and triplet geometries are in reality tautomeric forms. This
observation can be made from the geometry optimization at UHF/6-311G(d),
ROHF/6-311G(d), and ROB3LYP/6-311G(d,p) levels.

Table 3.3. The ab-initio total energy and the optimized geometry for the spin states of 8-
imino-1-naphthalenyl methyl (3).

ES− ET
Method Basis sets Optimization ET(a.u.) <S2>T
(kcal mol−1)

UHF 6-311G(d) O –476.3133 3.1359 63.3


UB3LYP 6-311G(d) SP –479.3636 2.0931 24.3

SROHF 6-311G(d) O −476.2455 2.0000 20.8


ROHF 6-311G(d,p) O −476.2638 2.0000 19.9
ROB3LYP 6-311G(d,p) O −479.3741 2.0000 16.4

3.3.4. 1,8−naphthalenediylbis(amidogen)
The molecule 1,8−naphthalenediylbis(amidogen) (4) is a symmetric molecule with C2v
optimized geometry in each spin state. The triplet state is found to be the overall ground state
at each level of calculation. However when dealing with such systems one cannot rely fully

46
Chapter 3 Organic Diradicals

on nonpolarized bases. The molecule appears to be a prime candidate for DFT breakdown.
The nitrogen atoms of the diradical have three available electrons out of which two electrons
take part in π−bond formation while one electron is still left in a nonbonding orbital. These
NBMOs are degenerate, leading to difficulties in obtaining a consistent energy gap by B3LYP
or other DFT methods in the unrestricted formalism. DFT breakdown does not occur in the
restricted (open-shell) calculations. The best S−T value is 8.7 kcal mol−1 that has been
obtained from the R(O)B3LYP/6-311G(d,p) calculation [Table 3.4].

Table 3.4. The ab-initio total energy and the optimized geometry for the spin states of
1,8- naphthalenediylbis(amidogen) (4) in the unrestricted Hartree-Fock
formalism.

ES− ET
Method Basis sets Optimization ET(a.u.) <S2>T
(kcal mol−1)

UHF 6-311G(d) O −492.3102 3.1644 80.8


UB3LYP 6-311G(d) SP –495.3982 2.1019 17.3

ROHF 6-311G(d) O –492.2395 2.0000 36.4


ROHF 6-311G(d,p) O –492.2577 2.0000 36.2
ROB3LYP 6-311G(d,p) O –495.4077 2.0000 8.7

3.3.5. 8-methyl-1-naphthyl carbene


The species 8-methyl-1-naphthyl carbene (5) has the point group C1 but the
framework has Cs symmetry. The species can be derived from molecule 2 by [1,8] migration
of a hydrogen atom so that it can be viewed as a slightly higher-energy isomer of molecule 2.
It is found to have a triplet ground state, in agreement with the prediction of Platz et al.4
Unlike the earlier species, the diradical center lies on a single atom. Table 3.5 shows the
computed energy gap between Singlet and Triplet states. The best value for the energy
difference is 7.4 kcal mol−1 [R(O)B3LYP/6-311G(d,p)]. The post-Hartree-Fock CC
calculations involving the split-valence bases were performed with the orbitals 27-124 active
in the CC expansion. However, for the minimal basis, the CC expansion apparently stabilizes
the singlet to a greater extent than the triplet, thus reducing the S-T energy gap drastically.
This happens whenever the diradical is centered on a single atom, that is, also for molecule 6.
47
Chapter 3 Organic Diradicals

In such cases, the CC calculation on the triplet using the minimal basis set is not at par with
the Singlet state calculation. The DFT calculations exhibit a systematic trend.
The S−T gap for this molecule is found to be generally higher than that for molecule 2
in the unrestricted formalism and lower in the restricted (open-shell) calculations. The
NBMOs are nondegenerate, and DFT gives rise to fairly good energy differences.

Table 3.5. The ab-initio total energy and the optimized geometry for the spin states of
8−methyl−1−naphthyl carbene (5) in the unrestricted Hartree-Fock formalism.

ES− ET
Method Basis sets Optimization ET(a.u.) <S2>T
(kcal mol−1)

UHF 6-311G(d) O −460.2737 3.1921 45.2


UB3LYP 6-311G(d) SP −463.2758 2.0385 8.0

ROHF 6-311G(d) O −460.2330 2.0000 22.3


ROHF 6-311G(d,p) O –460.2504 2.0000 19.8
ROB3LYP 6-311G(d,p) O –463.2940 2.0000 7.4

Table 3.6. The ab-initio total energy and the optimized geometry for the spin states of 8-
methyl-1-naphthalenyl imidogen (6) in the unrestricted Hartree-Fock
formalism. S and T indicate singlet and triplet respectively.

ES− ET
Method Basis sets Optimization ET(a.u.) <S2>T
(kcal mol−1)

UHF 6-311G(d) O –476.3248 3.2048 70.5


UB3LYP 6-311G(d) SP −479.3681 2.0641 32.2

ROHF 6-311G(d) O −476.2719 2.0000 37.4


ROHF 6-311G(d,p) O −476.2923 2.0000 40.7
ROB3LYP 6-311G(d,p) O −479.3773 2.0000 27.2

3.3.6. 8−methyl−1−naphthalenyl imidogen


The molecule 8−methyl−1−naphthalenyl imidogen (6) has a planar framework in both
singlet and triplet states with symmetry Cs, the overall point group being C1. Table 3.6 shows
the computed energy gap between singlet and triplet states. From all levels of calculation the
triplet state emerges as the ground state. A single atom is the diradical centre, and again we
observe that the computed S−T energy gaps from CC calculations are unreliable while using
small bases. The species can be viewed as an analog of molecule 5, but the computed energy

48
Chapter 3 Organic Diradicals

gaps are a lot different from those of 5 because of the presence of a heteroatom. It can also be
considered as a higher-energy isomer of molecule 3. The best calculated S−T gap is 27.2 kcal
mol−1 [R(O)B3LYP/6-311G(d,p)].
Like molecule 5, species 6 has almost the same singlet and triplet geometries. The
singlet state has slightly elongated bonds compared to the triplet state, with the differences
varying upto 0.15Å. The bond angles hardly vary in the two spin states.

3.3.7. 8−methyl−1−naphthyl diazomethane


The molecule 8−methyl−1−naphthyl diazomethane (7) shows a large difference in its
singlet and triplet optimized structures. The singlet has Cs symmetry while the point group
for the triplet is C1. See Table 3.7. The NBMOs in this case are somewhat degenerate
leading to a slight DFT breakdown. Because of the presence of the bonds between the
heteroatoms, the UHF calculation using the STO-3G basis very confusingly indicates the
triplet as the ground state. But this is corrected by using the split-valence basis which shows
the Singlet and Triplet to have almost the same energy. The post−Hartree−Fock calculations
invariably point out the singlet to be the more stable species. Here again, the density
functional treatment leads to a systematic trend in the S−T energy difference. Calculations
were also performed using the polarized basis sets to yield the best S−T gap as −21.9 kcal
mol−1 at the R(O)B3LYP/6-311G(d,p) level.

24
N 24 17
21 N 23
H 22 N H H17 18
H H H
22
H N H18 21
20 16
20 23 16 1 15
H 19
1 15 3 H
H 19 H 13
14
3 14 2
2 13 4
H 6
8 11 H12
H 4 8 11 H12 5 9
6 9
5 H H 10
H H 7
7 10
S T
Figure 3.3. Optimized Singlet (S) and Triplet (T) structures for molecule 7.
49
Chapter 3 Organic Diradicals

In the singlet state, the molecule takes up a very interesting geometry. One of the
nitrogen atoms becomes equidistant from the two CH2 groups with the length of the C−N bond
of the same order as that of a carbon−nitrogen single bond (Figure 3.3). This points out the
formation of a stable six membered ring by one nitrogen and five carbon atoms. This happen
in all the calculations, both restricted and unrestricted. The nitrogen atoms, however, remain
out of the plane of the carbon atoms so that the point group is only Cs. This situation does not
hold for the triplet case where the nitrogen atom points away from the carbon atom C16 to
which it is not directly bonded.

Table 3.7. The ab-initio total energy and the optimized geometry for the spin states of 8-
methyl-1- naphthyl diazomethane (7) in the unrestricted Hartree-Fock
formalism.

ES− ET
Method Basis sets Optimization ET(a.u.) <S2>T
(kcal mol−1)

UHF 6-311G(d) O −569.1994 3.3101 12.9


UB3LYP 6-311G(d) SP −572.8201 2.0563 −17.9

ROHF 6-311G(d) O −569.1501 2.0000 −18.0


ROHF 6-311G(d,p) O −569.1676 2.0000 −17.3
ROB3LYP 6-311G(d,p) O −572.8337 2.0000 −21.9

3.4. Conclusions
The diradicals TMM, TME and m-xylylene were investigated earlier by a large
12−14
number of researchers in great detail. In a previous occasion, Datta et al.6 discussed the
spin nature of some chain and monocyclic diradicals, and found the UCCSD(T) methodology
with split-valence basis sets to be a dependable approach to the calculation of the S−T energy
difference for diradicals. Here we have explored the ground electronic spin state of seven
organic diradicals belonging to the condensed ring system.
Though the UHF method gives a more or less correct, optimization of the molecular
geometry in each spin state, the relative energies calculated by the UHF method are not
reliable. Hence the method often yields significantly wrong S−T energy differences. The
calculated S−T splitting vastly improves by the application of coupled-cluster methods on the
UHF optimized geometry. The other alternative, Møller-Plesset perturbation theory,

50
Chapter 3 Organic Diradicals

generally yields misleading results for the S−T gap. This is also in general agreement with
the observations of Mitani et al.14g,h Results from the MP calculations are not shown in this
report. The S−T energy gap calculated with small basis sets like STO−3G and 4−31G at the
UCCSD(T) level is not very realistic. One has to use larger basis sets, especially those with
polarization functions. This imposes a limit on the computing ability using coupled-cluster
methods. So, one can resort to the density functional treatment as a workable solution.
There is another aspect of the problem. The unrestricted formalism gives rise to a
very high spin contamination as can be seen from the S2 expectation value computed for the
triplet state. The post-Hartree-Fock methods do not significantly rectify this error. But the
density functional treatment such as UB3LYP reduces spin-contamination and yields <S2>T of
the order of 2.1. The spin-contamination effect is best avoided by the restricted formalism.
The RHF formalism suffers from the difficulty that a much larger basis set is needed to obtain
the correct triplet geometry. This is why we carried out single point UB3LYP calculations
using 6-311G(d) basis sets whereas for each restricted calculations the geometry was
explicitly optimized.

Tg S

Figure 3.4. The PM3 spin density contours for molecules 1 in Singlet (S) and Triplet (T) states.
The superscript g indicates the calculated ground state. Plots for 2-7 can be found in
ref 20.

The DFT methodology does not always work in the unrestricted formalism. In fact,
we have noticed more or less a systematic trend in the UB3LYP calculations only for
molecules 5-7. Our best results are, therefore, from the R(O)B3LYP/6-311G(d,p)
51
Chapter 3 Organic Diradicals

calculations. The experimental gaps, once they are measured, are predicted to be found
within a few kcal mol−1 of the values calculated here. The calculated spin ordering in the
ground states are in excellent agreement with the experimental observations discussed in
section 1.4,7-10,12−14
All the molecules except species 7 have triplet ground states. Molecule 7 is the only
species investigated here in which one of the diradical centers is not directly attached to the
ring. This indicates a very low S−T gap or even a ground state singlet. This prediction is
borne out by all the calculations except the UHF ones.
The optimized structures in the singlet and triplet states vary from each other.
Symmetry breaking has been found to be essential in obtaining a correct estimate of the S−T
splitting that is usually of the order of only a few kcal mol−1. The variation has been found to
be the largest for molecules 3 and 7. The structure of molecule 3, as shown in Figure 3.2, is
representative of the stable, that is, the triplet state. The stable singlet is a tautomeric form
with substituents −NH2 and −CH in lieu of −NH and −CH2. The singlet of molecule 7 has a
three-fused-ring non-planar structure that has been evidenced by geometry optimization by all
the methods indicated in Table 3.7. See Figure 3.3. Finally, the rule of spin alternation in
UHF is again found to be robust. It can be used to identify the correct spin nature of the
ground state without fail for the diradical systems as shown in Figure 3.4.

3.5. References

1. (a) Coulson, C. A. J. Chim. Phys. 1948, 45, 243. (b) Moffitt, W. E. J. Chem. Soc., Faraday Trans. 1949,
45, 373.
2. Borden, W. T.; Davidson, E. R. J. Am. Chem. Soc. 1977, 99, 4587.
3. Hoffmann, R. J. Am. Chem. Soc. 1968, 90, 1475.
4. Platz, M. S.; Carrol, G.; Pierrat, F.; Zayas, J.; Auster, S. Tetrahedron, 1982, 38, 777.
5. (a) Conrad, M. P.; Pitzer, R. M.; Schaefer III, H. F. J. Am. Chem. Soc. 1979, 101, 2245. (b) Borden, W.
T.; Davidson, E. Annu. Rev. Phys. Chem. 1979, 30, 125.
6. Datta, S. N.; Mukherjee, P; Jha, P. P. J. Phys. Chem. A. 2003, 107, 5049.
7. Rule, M.; Matlin, A. R.; Hilinski, E. F.; Dougherty, D. A.; Berson, J. A. J. Am. Chem. Soc. 1979, 101,
5098.
8. Pagni, R. M.; Burnett, M. N.; Dodd, J. R. J. Am. Chem. Soc. 1977, 99, 1972.
9. Trozzolo, A. M.; Wasserman, E.; Yager, W. A. J. Am. Chem. Soc. 1965, 87, 129.
10. Platz, M. S.; Burns, J. R. J. Am. Chem. Soc. 1979, 101, 4425.

52
Chapter 3 Organic Diradicals

11. Borden, W. T.; Davidson, E. R.; Feller, D. Tetrahedron 1982, 38, 737.
12. (a) Cramer, C. J.; Smith, B. A. J. Phys. Chem. 1996, 100, 9664. (b) Feller, D.; Davidson, E. R.; Borden,
W. T. Isr. J. Chem. 1983, 23, 105. (c) Dietz, F.; Schleitzer, A.; Vogel, H.; Tyutyulkov, N. Z. Phys.
Chem.(Munich), 1999, 209, 67. (d) Gisin, M.; Wirz, J. Helv. Chim. Acta. 1983, 66, 1556. (e) Lahti, P.
M.; Rossi, A. R.; Berson, J. A. J. Am. Chem. Soc. 1985, 107, 2273. (f) Prasad, B. L. V.; Radhakrishnan,
T. P. J. Phys. Chem. 1992, 96, 9232. (g) Li, S.; Ma, J.; Jiang, Y. J. Phys. Chem. A. 1997, 101, 5587. (h)
Li, X.; Paldus, J. Chem. Phys. 1996, 204, 447. (i) Li, S.; Ma, J.; Jiang, Y. J. Phys. Chem. A. 1997, 101,
5567. (j) Pranata, J. J. Am. Chem. Soc. 1992, 114, 10537. (k) Shen, M.; Sinanoglu, O. Stud. Phys.
Theor. Chem. 1987, 51, 373.
13. (a) Hashimoto, K.; Fukutome, H. Bull. Chem. Soc. Jpn. 1981, 54, 3651. (b) Du, P.; Borden, W. T. J.
Am. Chem. Soc. 1987, 109, 930. (c) Nachtigall, P.; Jordan, K. D. J. Am. Chem. Soc. 1992, 114, 4743.
(d) Nachtigall, P.; Jordan, K. D. Ibid. 1993, 115, 270. (e) Rodriguez, E.; Reguero, M.; Caballol, R. J.
Phys. Chem. A, 2000, 104, 6253. (f) Filatov, M.; Shaik, S. Ibid. 1999, 103, 8885. (g) Chakrabarti, A.;
Albert, I. D. L.; Ramasesha, S.; Lalitha, S.; Chandrasekhar, J. Proc. Ind. Acad. Sci. 1993, 105, 53. (h)
Prasad, B. L. V.; Radhakrishnan, T. P. THEOCHEM, 1996, 361, 175. (i) Mahlmann, J.; Kleissinger, M.
Int. J. Quantum Chem. 2000, 77, 446. (j) Pittner, J.; Nachtigall, P.; Carsky, P.; Hubac, I. J. Phys. Chem.
A, 2001, 105, 1354.
14. (a) Kato, S.; Morokuma, K.; Feller, D.; Davidson, E. R.; Borden, W. T. J. Am. Chem. Soc. 1983, 105,
1791. (b) Karafiloglou, P. Croat. Chem. Acta, 1983, 56, 389. (c) Karafiloglou, P. Int. J. Quantum Chem.
1984, 25, 293. (d) Fort, Jr. R. C.; Getty, S. J.; Hrovat, D. A.; Lahti, P. M.; Borden, W. T. J. Am. Chem.
Soc. 1992, 114, 7549. (e) Fang, S.; Lee, M. S.; Hrovat, D. A.; Borden, W. T. J. Am. Chem. Soc. 1995,
117, 6727. (f) Baumgarten, M.; Zhang, J.; Okada, K.; Tyutyulkov, N. Mol. Cryst. Liq. Sci. Technol.
Sect. A. 1997, 305, 509. (g) Mitani, M.; Mori, H.; Takano, Y.; Yamaki, D.; Yoshioka, Y.; Yamaguchi,
K. J. Chem. Phys. 2000, 113, 4035. (h) Mitani, M.; Yamaki, D.; Takano, Y.; Kitagawa, Y.; Yoshioka,
Y.; Yamaguchi, K. J. Chem. Phys. 2000, 113, 10486. (i) Lejeune, V.; Berthier, G.; Despres, A.;
Migirdicyan, E. J. Phys. Chem. 1991, 95, 3895. (j) Sandberg, K. A.; Shultz, D. A. J. Phys. Org. Chem.
1998, 11, 819. (k) Havlas, Z.; Michl, J. J. Chem. Soc. Perkin Trans. 2, 1999, 11, 2299.
15. (a) Klein, D. J.; March, N. H. Int. J. Quantum Chem. 2001, 85, 327. (b) Ivanciuc, O.; Bytautas, L.;
Klein, D. J. J. Chem. Phys. 2002, 116, 4735. (c) Ivanciuc, O.; Klein, D. J.; Bytautas, L. Carbon 40,
2002, 2063.
16. Hrovat, D. A.; Murcko, M. A.; Lahti, P. M.; Borden, W. T. J. Chem. Soc., Perkin Trans. 2, 1998, 5,
1037.
17. Kahn, M. I.; Goodman, J. L. J. Am. Chem.. Soc. 1994, 116, 10342.
18. HyperChem Professional Release 7 for Windows, (Hypercube Inc., Gaimesville, 2002).
19. Frisch, M. J.; et al. Gaussian 98; Gaussian, Inc.: Pittsburgh, PA, 1998. Gaussian 98 for Windows,
(Gaussian Inc., Pittsburgh, 2002).
20. Jha, P.P.; Ali, Md. E.; Datta, S. N. J. Phys. Chem. A 2004, 108, 4084.
53
Chapter 4
Bis-Nitronyl Nitroxide Diradicals: Influence of Length and
Aromaticity of Couplers

A series of Nitronyl Nitroxide (NN) diradicals with linear conjugated couplers and another
series with aromatic couplers have been investigated by broken-symmetry (BS) DFT
approach. The overlap integral between the magnetically active orbitals in the BS state has
been explicitly computed and used for the evaluation of the magnetic exchange coupling
constant (J). The calculated J values are in very good agreement with the observed values in
literature. The magnitude of J depends on the length of the coupler as well as the
conformation of the radical units. The aromaticity of the spacer decreases the strength of the
exchange coupling constant. The SOMO-SOMO energy splitting analysis where SOMO
stands for the singly-occupied molecular orbital, and the calculation of electron paramagnetic
resonance (EPR) parameters have also been carried out. The computed hyperfine coupling
constants support the intramolecular magnetic interactions. The nature of magnetic exchange
coupling constant can also be predicted from the shape of the SOMOs as well as the spin
alternation rule in the unrestricted Hartree-Fock (UHF) treatment. It is found that π-
conjugation along with the spin-polarization plays the major role in controlling the magnitude
and sign of the coupling constant.
Chapter 4 Nitronyl Nitroxide

4.1. Introduction
Nitronyl nitroxide is found to be one of the most promising radicals in molecular
magnetism due to its exceptional stability, facile method of preparation, versatility in
coordination, and ability to generate cooperative magnetic properties. Here we report the
results of the study of a series of nitronyl nitroxide based diradicals with different conjugated
magnetic couplers. In the present work we establish that the strength of the magnetic
interaction decreases with the increase in size of the conjugated coupler in a quantitative way,
and also with the extent of aromaticity of the ring coupler. With this aim, we have studied a
series of NN diradicals with different magnetic couplers: No coupler (1), the ethylenic coupler
(2), 1,4 butadienic coupler (3), 1,6-hexatrienic coupler (4), p-phenylene coupler (5), 2,6-
pyridine coupler (6), m-phenyelene coupler (7), 2,5-furanic coupler (8), 2,5-pyrrolic coupler
(9), and 2,5 m-thiopheneic coupler (10) . All the couplers are π-conjugated molecules.
Three recently synthesized diradicals, (11-13) with larger linear diphenylene acetylene
couplers are also investigated to study the dependence of J on basis sets.

4.2. Theoretical Background


In this work we have explicitly computed the overlap integral Sab . The α-HOMO and

β-HOMO in the BS state have been considered as the magnetic orbitals. It is observed that
the overlap between the magnetic orbitals is very low for all the diradicals except 1a, 1c and
1d. We have further noticed that the <S2> value for all the calculated BS states deviate very

little from 1.00, and, in particular, the difference ( < S 2 >T − < S 2 > BS ) is nearly equal to
unity for these systems. Therefore, the magnetic exchange constants have been calculated
here by using both Eq. (2.2) and Eq. (2.4). Only in the moderately large Sab region, for 1a,

1c, 1d and 2, Eq.(2.2) is estimated to yield better result.

4.3. Computational Strategy


The molecular structures of all the diradicals 1-10 (Figure 4.1) have been fully
optimized at ROHF/6-31G(d,p) level. The optimized dihedral angle of diradical 1 between
the two planes of the NN moiety has been found to be 78° in the isolated molecule. But the
crystallographic data suggests that the dihedral angle is 55°.1 So we have taken several values

54
Chapter 4 Nitronyl Nitroxide

of the dihedral angle between the two NN moieties while keeping the rest of the optimized
molecule intact, and computed the exchange coupling constant for each of these geometries.
The angles considered are 0°, 55°, 78° and 90°.
Single point calculations have been performed on the optimized geometry at the
UB3LYP level with 6-311G(d,p) and 6-311+G(d,p) basis sets. To obtain the broken-
symmetry states, single-point UB3LYP calculations have been carried out using the accurate
guess values of molecular orbitals, which are in turn retrieved from the proper ROHF
calculations. These calculations have been done by using Gaussian 98 quantum chemical
package. The visualization software Molden2 and Molekel3 have also been used. The overlap
integral between the two magnetic orbitals in the BS state has been calculated by a program of
our own. This program utilises the MO coefficients and basis set information at 6-311+G(d,p)
level from the Gaussian 98 log files.
To further support the magnetic properties, the hyperfine coupling constants (hfcc)
have been calculated at B3LYP level by using EPR-II and EPR-III basis sets.4 The diradical
10 contains one S atom, but this atom is not included in the EPR basis set of Gaussian 98.
Therefore, during the calculation of hfcc we have used 6-311G(d,p) basis set for the S atom,
while EPR-II and EPR-III basis sets have been used for the rest of the atoms.

4.4. Results and Discussion


First of all, to make the discussion clear, the computed overlap integrals ( Sab ) are

given in Table 4.1. The moderately large overlap region is manifest for 1a, 1c, 1d and 2. For
these species, neither Eq. (2.4) nor Eq. (2.5) can be used with accuracy. Therefore, Eq. (2.2)
gives a better estimate of J value. For 1b and 9, Eq. (2.2) would make a deviation of about
3% and 2% respectively from the J value calculated from Eq. (2.4). For all others, Eq. (2.4)
represents a better choice.
The calculation of the intramolecular exchange coupling constant between the two NN
monoradicals without any coupler (in species 1) is shown in Table 4.2. The J values for the
planar diradicals with no coupler and π-conjugated linear couplers are tabulated in Table 4.3.
The values for the six member and five member conjugated aromatic couplers are given in
Table 4.4 and Table 4.5 respectively.

55
Chapter 4 Nitronyl Nitroxide

Table 4.2 shows that the magnitude of J drastically decreases with the increase in the
dihedral angle. The highest J is −923 cm−1 for the planar configuration and lowest value is
1 6 6
Ο Ο Ο
+ 1 +
2
N 7N Ο 7N
3 8 2 12 8
N
4 + 3 11
N 9N 9N
4+
5Ο 10 N 10
Ο Ο
5
Ο
1 2

6Ο 1 6Ο
2 Ο
1 +
2 Ο 7N
+
N 14 16 7N
14 12 8
N 12 8
4+
3
4+ 3 N 11 13 15 9N
13 9N
N 11 5 10
10 Ο Ο
5
Ο Ο 4
3

1 13
Ο 6Ο 1
2 12 + Ο 12 14 6Ο
13 7N
N 2N +
3 11 14 8 3 8 7N
11 N 15
4 +
N 9N 4 + 16 9N
16 15 N
5Ο 10 10
Ο 5Ο Ο
5 6

13 1
1Ο 6Ο Ο 12 13 6Ο
12 14
+ 2N 11 14 7 +
2N 3 8 7N 3 8 N
11 15 O15
+
16 4 + 9N
4
N 9N N 10Ο
10
Ο 5Ο

7 8

1Ο 12 13 6Ο 1Ο 12 13 6Ο
2N 11 14 7 +
2N 11 14 7 +
3 8 N 3 8 N
N15 S15
4 + H 9N 4 + 9N
N N
10Ο 10
5Ο 5Ο Ο

9 10

Figure 4.1. The systems under investigation with (1) no coupler, (2) ethylenic coupler, (3) 1,4-
butadienic coupler, (4) 1,6-hexatrienic coupler, (5) p-phenylene coupler, (6) 2,6-
56
Chapter 4 Nitronyl Nitroxide

pyridinic coupler, (7) m-phenylene coupler, (8) 2,5-furanic coupler, (9) 2,5-pyrrolic
coupler, and (10) 2,5-thiophenic coupler between the two nitronyl nitroxide
monoradicals.

Table 4.1. The computed overlap integral between the two magnetically active orbitals in Broken-
symmetry state. The computed results are for the 6-311+G(d,p) basis sets.

Sl. No Coupler Sab

1a 0° dihedral angle −0.494041


1b 55° dihedral angle −0.178056
1c 78° dihedral angle −0.791540
1d 90° dihedral angle 0.569932

2 ethelenic 0.361410
3 1,4-butadienic 0.039911
4 1,6-hexatrinenic −0.072483

5 p-phenylene −0.014067
6 m-pyridinic 0.044348
7 m- phenylene −0.051216

8 2,5-furanic −0.006970
9 2,5-pyrrolic 0.134857
10 2,5-thiophenic 0.036758

−29 cm−1 for the 90° rotated species. This is due to the maximum overlap between the two p-
orbitals in bridging carbon atoms when the dihedral angle is 0°, and the minimum conjugation
when the two p-orbitals in bridging atoms are orthogonal. In crystal structure of 1 it is
observed that the dihedral angle is 55°. The J value calculated for 1b by using Eq. (2.4)
excellently matches with the observed J in molecular crystals. The trend in Table 4.2 makes
it amply clear that the delocalization of the π-electrons plays the major role in controlling the
exchange coupling constant. The larger dihedral angle inhibits conjugation of the π-electrons.
Nevertheless, a weak antiferromagnetic interaction exists even when the two p-orbitals are
orthogonal to each other. In this case, there is a strong localization of the SOMOs. The spin
of the unpaired electron in one of the π-orbitals polarizes the spin of the paired electrons in
the orthogonal σ-orbital. The residual spin polarization is the sole reason for a very weak
antiferromagnetic coupling constant in 1d.
It is observed that the exchange coupling constant decreases with the increase of the
length of the coupler (Table 4.3). In this Table, Eq. (2.2) is a better description for 1a while
Eq. (2.4) is more appropriate for 25a, 3 and 4.6 This is a very normal trend. It is observed that
57
Chapter 4 Nitronyl Nitroxide

2 has the highest J value. Our theoretical calculations have also supported this finding.6 The
main reason for it is that the steric effects force the dihedral angle of 1 to be 55° in molecular
crystal, which causes loss of delocalization. The rule of spin alternation in the UHF treatment7
can also predict the proper ground spin state for all the cases in Table 4.2 and Table 4.3
(Figure 4.1).

Table 4.2. Single-point energies and calculated intramolecular exchange coupling constants for
the Nitronyl Nitroxide (NN) diradicals without any coupler. The coupling constant J is
calculated for different dihedral angles. All the single-point calculations are
performed with the UB3LYP methodology for the broken-symmetry state as well as
the triplet state.

Dihedral Basis Energy (a.u.) J


angle sets <S2> (cm−1)
BS T Calculated Exptl.
Eq. (2.2) Eq. (2.4)

6-311G** −1067.7727349 −1067.7675028 −923 −1148


0° 1.125262 2.046992 NAa
1a 6-311+G** −1067.7904005 −1067.7851668 −923 −1148
1.111334 2.046176

6-311G** −1067.8615782 −1067.8602569 −281 −290


55° b 1.077337 2.060998 −311c
1b 6-311+G** −1067.8806805 −1067.8793779 −277 −286
1.07538 2.06042

6-311G** −1067.864504 −1067.8642504 −34 −56


78° d 1.073739 2.066934 NAa
1c 6-311+G** −1067.8848812 −1067.8845819 −41 −66
1.070904 2.064865

6-311G** −1067.8637813 −1067.8636461 −23 −30


90° 1.072735 2.068682 NAa
1d 6-311+G** −1067.884201 −1067.8840684 −22 −29
1.070115 2.066156

a
Not available in literature; b Rotating the N-C-C-N dihedral angle of fully optimized geometry to 55° so as to
get a structure similar to the crystallographic one; c Ref. 1; d Fully optimized geometry at ROHF/6-311G(d,p)
level.

The calculated J values are in very good agreement with the observed values for 5-7 in
Table 4.4. Here, we find hardly any difference between Eqs. (2.2) and (2.4). The length of
the coupler in 5 is similar to the butadienic coupler in 3. However, the magnetic exchange
coupling constant is found to be less than that for the linear conjugated coupler. In general,

58
Chapter 4 Nitronyl Nitroxide

all the conjugated aromatic couplers are weaker than the liner couplers. The spin alternation
rule for the prediction of the ground state spin is also supported by the experimental results on
1,4 phenylene (5),8 2,6-pyiridinic coupler (6)9 and 2,6 phenylene coupler (7),10 with singlet,
triplet and triplet ground states respectively.

Ο Ο Ο Ο
+ +
N N N N
(CH=CH )n +
+
N N N N
Ο Ο Ο Ο

1, 2, 3, 4 for n = 0, 1, 2, 3 5
Antiferromagnetic Antiferromagnetic

Ο Ο Ο Ο
+ +
N N N N
X X
+ +
N N N N
Ο Ο Ο Ο

6, 7 for X = N, C 8, 9, 10 for X = O, NH, S


Ferromagnetic Antiferromagnetic

Figure 4.2. Prediction of ground spin states and hence the nature of the magnetic exchange
coupling constants are shown according to the spin alternation rule.

Results for 8, 9 and 10 are given in Table 4.5. Here, again, the GND expression (2.4)
gives a more reliable estimate of J in every case. The data for 5 has been included in this
table for the reason of making a facile comparison. The calculated J is in good agreement
with the observed value for 10.11 Experimental values are lacking for 8 and 9, and the J values
−148 cm-1 and −164 cm−1 are predicted estimates. Again, the spin alternation rule identifies
the proper ground state for 10 as a singlet. The identified ground states for 8 and 9 are both
singlet, in agreement with the computed J values.
The sign of J depends on the parity of the number of bonds in the coupling pathway
through the coupler. When the number of bonds is odd, J is negative like in 1a, 2, 3 and 4 (1,
3, 5 and 7 bonds). In 5, there are two five-bond coupling pathways (odd number) and the
resulting J value is negative. In 6 and 7, there are two even coupling pathways (four- and six-
59
Chapter 4 Nitronyl Nitroxide

bond couplings), and J is positive. These observations represent a mere restatement of the so-
called spin alternation rule (Figure 4.2). In all three cases 8, 9, and 10, there are one even
Table 4.3. Single-point energies and calculated intramolecular exchange coupling constants for π-
conjugated linear couplers. All the single-point calculations are performed with the
UB3LYP methodology for the broken-symmetry state as well as the triplet state.

Energy (a.u.) J
Diradical Basis sets <S2> (cm−1)
BS T Calculated Exptl
.
Eq. Eq.
(2.2) (2.4)

6-311G** −1067.7727349 −1067.7675028 −923 −1148


NN NN 1.125262 2.046992 NAb
1a 6-311+G** −1067.7904005 −1067.7851668 −923 −1148
1.111334 2.046176

6-311G** −1145.3113872 −1145.3096214 −343 −388


NN 1.139187 2.066011 –350c
NN 6-311+G** –1145.3287469 –1145.3271496 −310 −350
2 1.1286 2.0629

6-311G** –1222.7385572 –1222.7374101 −251 −251


NN 1.144146 2.084033 NAb
NN 6-311+G** –1222.7589119 –1222.7578636 −230 −230
3 1.134768 2.080213

NN 6-311G** −1300.1675127 −1300.1668185 −151 −152


NN 1.130756 2.083696 −66d
4 6-311+G** −1300.1870287 −1300.1864071 −135 −136
1.120468 2.078546

a
Rotating the N-C-C-N dihedral angle of fully optimized geometry to 0° so as to get a planar structure like 2 and
3; b Not available in literature; c Ref. 5(a); d Ref. 5(b); for 1,6 dimethyl derivative.

and one odd pathway. At a first glance, one would think that there is a competition between
the two pathways. In reality, the odd (five-bond) route is supported by the even (four-bond)
path through the heteroatom as the latter contributes two π electrons. The J values for 8, 9,
and 10 are all negative. In magnitude, these are actually larger than the J value for 5 (Table
4.5). This behavior is similar to that known for the Fermi Contact contribution to nuclear
spin-spin couplings transmitted through the π-electronic system in conjugated compounds,
and can be viewed as an extension of the spin alternation rule to the case of heteronuclear

60
Chapter 4 Nitronyl Nitroxide

aromatic couplers.

Table 4.4. Single-point energies and calculated intramolecular exchange coupling constants for
aromatic couplers. All the single-point calculations are performed with the UB3LYP
methodology for the broken-symmetry state as well as the triplet state.

Diradicals Basis Energy (a.u.) J


sets <S2> (cm−1)
BS T Calculated Exptl.
Eq. (2.2) Eq. (2.4)

6-311G** −1298.9901966 −1298.9897665 −94 −94


NN
1.090979 2.073425 −72a
NN
6-311+G** −1299.0106241 −1299.0102268 −87 −87
5 1.086835 2.070655

6-311G** −1315.0163354 −1315.0164212 19 19


1.069793 2.074852 7b
6-311+G** −1315.0289297 −1315.0290336 23 23
NN N NN 1 .083623 2.090128
6

6-311G** −1298.9864545 −1298.9865576 23 23


1.073124 2.078921 20c
NN NN 6-311+G** −1299.0066208 −1299.00 6716 21 21
7 1.070268 2.075584

a
Ref 8, bRef 9, cRef 10.

4.4.1. Rationalization
The spin density distribution in all the species investigated here is more or less
(pairwise) symmetric for rotation by 180° around the principal axis (C2). An understanding of
the trend of the J values in each series can be obtained by writing

J = ∑ J ij ρi ρ j (4.1)
i= j

where ρi is the spin density on the ith atom in the triplet state, and Jij is the exchange integral

between the π-orbitals of the atoms i and j. The integral Jij is strongest for atoms i and j being
nearest neighbors. For a conjugated coupler of N atoms, there are (N+1) nearest neighbors.
But the absolute magnitude of the atomic spin density approximately varies as 1/(N+1).
Therefore, as N increases, the absolute magnitude of J decreases approximately as 1/(N+1).
This is a general trend, and Table 4.3 bears a glowing testimony to it. The trend is clearly set
61
Chapter 4 Nitronyl Nitroxide

with |J| exhibiting the order 1a > 2 > 3> 4 in the approximate ratio 1:1/3:1/5:1/7, and the
longer the coupler is, the less antiferromagnetic interaction is there.

Table 4.5. Single-point energies and calculated intramolecular exchange coupling constants for
five-member aromatic couplers. The p-phenylene diradical is included here for the
purpose of making a comparison possible. All the single-point calculations are
performed with the UB3LYP methodology for the broken-symmetry state as well as
the triplet state.

Diradicals Basis Energy (a.u.) J


sets <S2> (cm−1)
BS T Calculated Exptl.
Eq.(2.2) Eq.(2.4)

6-311G** −1298.9901966 −1298.9897665 −94 −94


NN NN
1.090979 2.073425 −72a
6-311+G** −1299.0106241 −1299.0102268 −87 −87
5 1.086835 2.070655

6-311G** −1296.7615506 −1296.760815 −161 −161


1.098239 2.065873
NN O NN 6-311+G** −1296.7837161 −1296.7830414 −148 −148 NAb
8 1.092527 2.062898

6-311G** −1276.9299062 −1276.9291011 −174 −177


1.100927 2.07081
6-311+G** −1276.9503153 −1276.9495684 −161 −164 NAb
NN N NN 1.095613 2.067769
H
9

6-311G** –1619.7607667 −1619.7599146 −187 −187


1.108292 2.07297
6-311+G** −1619.7811119 −1619.7803369 −170 −170 −157c
NN S NN 1.101501 2.069521
10

a
Ref. 8; bNot available in literature; cRef. 11.

In the case of 6-membered ring aromatic couplers, the rule of spin alternation indicates
that an antiferromagnetic coupling exists for o-phenylene and p-phenylene or their
derivatives, and a ferromagnetic coupling exists for m-phenylene. For 5-membered ring
heteronuclear aromatic couplers, the 2,3 and 3,4 species are to be treated as the o-couplers,
and 2,5 species is a p-coupler while the 2,4 one acts as m-coupler, because the hetero atom in

62
Chapter 4 Nitronyl Nitroxide

position 1 provides two π-electrons.


The chain rule here suggests that J ∝ 2(N+1)/(2N+1)2 where 2N is the number of
conjugated atoms in the coupler. Thus, J ∝8/49 for 6-membered p-couplers whereas J ∝ 1/5
for the butadienic coupler. Therefore, the magnitude of the J value decreases by ring
formation. The atomic spin densities in the coupler decrease further due to resonance. So a
6-membered π-coupler has a considerably reduced |J | compared to the value for a linear chain
of 4 carbon atoms. This is turned out by the calculated values for the butadienic coupler 3
(−230 cm−1) in Table 4.3 and the p-phenylenic coupler 5 (−87 cm−1) in Table 4.4. For the 6-
membered m-couplers like 6 and 7, as aromaticity increases, the J value increases (Table 4.4).
Therefore, aromaticity favors the ferromagnetic trend.
The heteronuclear couplers are less aromatic. Therefore, by counting all the six π
electrons, the para coupling with heteronuclear aromatic spacers would entail, and J ∝ 1/5.
But the resonance decreases the atomic spin densities. These two factors lead to a J value that
is almost midway between the J for 3 (−230 cm−1) and the J for 5 (−87cm−1). See Table 4.5.
Here again, the decrease in aromaticity is accompanied by an increase in antiferromagnetic
coupling. This is evidenced from the trend 5 < 8 < 9 < 10 for the absolute magnitude of the
calculated J values given in the same table.

4.4.2. SOMO-SOMO Energy Level Splitting


Hoffmann12 provided a criterion based on the extended Hückel calculations on
benzyne and diradicals, which suggests that if the energy difference between the two SOMOs
(∆ESS) is less than 1.5 eV, the two nonbonding electrons will occupy different degenerate
orbitals with a parallel-spin configuration so as to minimize their electrostatic repulsion and
thereby leading to a triplet ground state. Constantinides et al.13 have investigated a series of
4nπ antiaromatic linear and angular poly-heteroacene molecules by B3LYP/6-31G(d) method
and found that singlet ground states result when ∆ESS > 1.3 eV. Zhang et al.14 have
calculated a series of m-phenylene-bridged diradicals to investigate the effect of substitution
on the S-T energy gaps and ground state multiplicity. They have calculated ∆ESS at
ROB3LYP/6-31G(d) level. The low spin ground state results even when ∆ESS is found to be
0.19 eV. Our calculation of ∆ESS for all the diradicals 1-10 by ROB3LYP/6-311+G(d,p)
method in Table 4.6 does not reveal much information about the ground state spin.

63
Chapter 4 Nitronyl Nitroxide

For the diradicals 1a-1d the SOMO-SOMO energy gap decreases as the dihedral angle
increases. Thus the magnitude of the J value decreases with the decrease of the SOMO-
SOMO energy gap. This trend is in agreement with the Hay-Thibeault-Hoffmann (HTH)
formula for the triplet-singlet energy difference15 in a dinuclear complex containing two
weakly interacting metal atoms. For species 1, the SOMOs are not degenerate even when the
p-orbitals are orthogonal to each other. A very weak antiferromagnetic interaction is observed
in species 1d. However, as Table 4.6 shows, the same formula does not hold for other
diradicals examined here: species 6 and 7 with relatively large SOMO-SOMO energy gaps are
known to have ferromagnetic coupling and our calculations also support this fact (Table 4.4),
while the others have much smaller gaps but are antiferromagnetically coupled.
The shape of all the SOMOs at ROB3LYP for diradicals 1−10 are given in Figure 4.3.
In general, two types of SOMOs are found, namely, disjoint (where no atoms are common)
and nondisjoint (with common atoms). All the diradicals except 6 and 7 are nondisjoint in
nature. The SOMOs of 2 and 9 seem to be apparently disjoint, but these are in reality
nondisjoint as observed from the molecular orbital coefficients. We find that for the type of
organic diradicals studied here, the ferromagnetic interaction arises when the shapes of the
SOMOs are disjoint in nature as in 6 and 7 (Figure 4.3).

Table 4.6. The energy levels of two SOMOs and their energy differences (∆ESS) at ROB3LYP/6-
311+G(d,p) level for the diradicals 1-10.

ES(1) ES(2) ∆ESS


Diradicals (a.u.) (a.u.) (eV)

1a -0.08364 -0.11228 0.7793


1b -0.09236 -0.09764 0.1437
1c -0.09701 -0.09850 0.0405
1d -0.09818 -0.09837 0.0052
2 -0.09405 -0.09550 0.0395
3 -0.09599 -0.09625 0.0071
4 -0.09647 -0.09662 0.0041
5 -0.09393 -0.09410 0.0046
6 -0.09189 -0.09590 0.1091
7 -0.09272 -0.09661 0.1059
8 -0.09656 -0.09680 0.0065
9 -0.09670 -0.09681 0.0030
10 -0.09709 -0.09764 0.0150

A similar point of interest arises. Borden and Davidson had argued that if Hückel
64
Chapter 4 Nitronyl Nitroxide

NBMO’s are not localized to disjoint groups of atoms, the triplet would lie below the
corresponding open-shell singlet at the SCF level. Our results contradict this observation, but
are in good agreement not only with experiment (Table 4.4) but also with the prediction from
the rule of spin alternation.

4.4.3. Isotropic Hyperfine Coupling Constant


From the experimental work it is observed that the hfcc of the two equivalent nitrogen
atoms in Nitronyl Nitroxide monoradicals with different substitutions at α-carbon atoms is in
the range of 7.00−7.81 G.16 The hfcc does not strongly depend on the nature of the
substitution at the α-position, but solvents play a dominant role. Hfcc values for diradicals
with conjugated couplers decrease to half of the values for the corresponding monoradicals.
The experimental values lie in the range of 3-4.5 G for diradicals with different couplers.17
Cirujeda et al.18 calculated the hfcc for several α-nitronyl aminoxyl radicals by
B3LYP method using EPR-II basis sets. They found similar hfcc for the monoradicals with
similar steric constraints between the two rings. This fact also supports that the spin density
distribution in the phenyl ring is not strongly dependent on the nature and position of
substituents. In our publications, the detailed discussions were given on this issue.6, 19
Table 4.7 shows that although the computed hfccs are different for the four nitrogen
atoms in diradicals 1-10, the average hfcc is reliably generated. Similar discrepancy for
different N atoms was also found by other authors. From the spin density distribution in the
triplet state it is observed that the calculated spin density is not homogeneously distributed
through the O-N-C-N-O bond, though the two N atoms are chemically equivalent. This fact
results in different values of hfcc, and is supported by the SOMO’s in Figure 4.3.

4.5. Diphenylene acetylene couplers: A study of size and basis sets effects
We have also investigated NN and imino nitroxide (IN) diradicals with extended
couplers. The following three spin couplers have been chosen (1) 2,2'-(1,2-ethynediyldi-4,1-
phenylene)bis[4,4,5,5-tetramethyl-4,5-dihydro-1 H-imidozolyl-oxyl] (IN-2p-IN, 11), (2) 2,2'-
(1,2-ethynediyldi-4,1 3,1-phenylene)bis[4,4,5,5-tetramethyl-4,5-dihydro-1 H-imidozolyl-
oxyl] (IN-pm-IN, 12) and (3) 2,2'-(1,2-ethynediyldi-4,1 3,1-phenylene)bis[4,4,5,5-
tetramethyl-4,5-dihydro-1 H-imidazole-1-oxyl-3-oxide] (NN-pm-NN, 13).

65
Chapter 4 Nitronyl Nitroxide

The molecular geometries are optimized at ROHF/6-31G(d) level and J is calculated

Table 4.7. The highperfine coupling constants (hfcc) calculated at B3LYP level with EPR-II and
EPR-III basis sets.

Diradicals Basis sets aN2 aN4 aN7 aN9 Observed


(G) (G) (G) (G) aN

1a EPR-II 4.11025 1.04162 1.03287 4.16915


EPR-III 4.23427 1.15317 1.14087 4.29395

1b EPR-II 4.31866 1.06099 1.05223 4.38357


EPR-III 4.47133 1.21081 1.20199 4.53641

1c EPR-II 4.46136 1.06774 1.07244 4.39365


EPR-III 4.67802 1.12935 1.13714 4.60911

1d EPR-II 4.59041 1.02385 1.03317 4.52136


EPR-III 4.73748 1.17743 1.18690 4.66801

2 EPR-II 4.52596 1.50349 1.34520 4.44834


EPR-III 4.17437 1.77937 1.49466 4.23153

3 EPR-II 4.35821 1.71624 1.71544 4.35549


EPR-III 4.43620 1.73844 1.73774 4.43366

4 EPR-II 4.54130 1.72502 1.72495 4.54584


EPR-III 4.50506 1.80138 1.80096 4.50906

5 EPR-II 3.38184 1.65462 1.65528 3.38509


EPR-III 3.50442 1.74658 1.74706 3.50721

6 EPR-II 4.48825 1.66194 1.66113 4.44737 3.3a


EPR-III 4.59373 1.66533 1.66506 4.55285

7 EPR-II 3.51368 1.49883 1.44803 4.62565


EPR-III 3.76404 1.84418 1.80126 4.88496

8 EPR-II 4.83269 1.62270 1.62222 4.83306


EPR-III 4.94187 1.65987 1.65902 4.94245

9 EPR-II 3.78055 1.48492 1.46878 4.86759


EPR-III 3.92236 1.66391 1.64693 5.01510

10 EPR-II 4.33239 1.50318 1.50422 4.30763 3.7b


EPR-III 4.51469 1.59648 1.59748 4.48928

a
Ref. 9, bRef. 11.

according to the procedure described in theoretical background section but using Eq. (2.4).
The ground spin state is also determined using CASSCF methodology.

66
Chapter 4 Nitronyl Nitroxide

Figure 4.3. Triplet SOMOs for all the diradicals 4−7, plotted at ROB3LYP/6-311+G(d,p) level.

The UB3LYP total energies for the BS and triplet states of IN-2p-IN (11) are given in
Table 4.8. The stability consistently increases with the basis size. We have considered up to
6-311G(d,p) basis set, as the triplet geometry was optimized at the 6-31G(d) level and a larger
basis would not necessarily generate a good value of the energy difference (EBS−ET). Both
67
Chapter 4 Nitronyl Nitroxide

the BS and T wave functions suffer from spin contamination effects, but the difference in
<S2> remains more or less equal to 1.0. The intramolecular magnetic exchange coupling
constant J, calculated from Eq. (2.4), shows a smooth trend for all the basis sets. Our best
result, computed with the 6-311G(d,p) basis, is J = −3.60 cm−1 that corresponds to the largest
basis in Table 4.8 and a very small (<S2>T − <S2>BS −1), 1.36×10−4. The experimental J is
−3.37 cm−1. The difference may be attributed to the solvent effect which is not considered in
our computations, the slight difference of (<S2>T − <S2>BS) from 1, and the constraint of
geometry optimization at the HF/6-31G(d) level. The magnetic coupling is manifestly
antiferromagnetic, as predicted by the simple spin alternation rule (Figure 4.4). This is also in
agreement with the observation that linear diradical derivatives of IN with other couplers have
singlet ground states.21

H H H H
O
. H H H
N

N N
N
.
O
N
33 H
N
N N
O . H H H H O . H H
H H

11 12
.
O
N
− H H H +
O N
N O−
+ H
N
.O H H H H
13

Figure 4.4. Spin alternation in the diradicals. IN-2p-IN (11) shows a singlet ground state whereas
IN-pm-IN (12) and NN-pm-NN (13) have triplet ground states.

Results form UB3LYP single-point calculations on IN-pm-IN (12) and NN-pm-NN


(13) are given in Table 4.9 and Table 4.10 respectively. The coupling between the radical
sites is manifestly ferromagnetic in each case, with positive J values. The calculated J values,
however, vary erratically with basis sets. This is unlike the antiferromagnetic case in Table
4.8. Therefore, the calculated J values are plotted against (Δ<S2> −1), where Δ<S2> = <S2>T
Table 4.8. Total energy from UB3LYP single point calculations on IN-2p-IN (11) in both broken-
symmetry (BS) and triplet (T) states using different basis sets.

68
Chapter 4 Nitronyl Nitroxide

Basis EBS(a.u.) ET(a.u.) EBS−ET (Δ<S2> −1) J (cm–1)a


<S2> <S2> (cal mol-1) x104

6-31G(d) −1455.5299574 −1455.5299586 0.75 3.31 0.26


1.022834 2.023165
6-31G(d,p) −1455.5768206 −1455.5768137 −4.33 −3.53 −1.52
1.022828 2.022475
6-31+G(d,p) −1455.6213633 −1455.6213548 −5.33 −5.96 −1.87
1.024742 2.024146
6-311G(d) −1455.84900831.0 −1455.8489953 –8.16 −8.10 −2.85
22876 2.022066
6-311G(d,p) −1455.8948799 −1455.8948635 −10.29 1.36 −3.60
1.022596 2.022732
Observedb −3.37

a
From Eq. (2.4). b Ref. 20.

Table 4.9. Total energy from UB3LYP single point calculations on IN-pm-IN (12) in both broken-
symmetry (BS) and triplet (T) states using different basis sets.

Basis EBS(a.u.) ET(a.u.) EBS−ET (Δ<S2> −1) J (cm–1)a


<S2> <S2> (cal mol-1) x104

6-31G(d) −1455.5310178 −1455.5310264 5.40 5.04 1.89


1.023946 2.02445
6-31G(d,p) −1455.5778669 −1455.5778774 6.60 6.99 2.32
1.024046 2.024745
6-31+G(d,p) −1455.6228021 −1455.6228052 1.95 2.39 0.68
1.026287 2.026526
6-311G(d) −1455.8502143 −1455.8502274 8.22 9.64 2.87
1.024004 2.024968
Extrapolatedb 0.16

a
From Eq. (2.4). b Fig. 4.5 (a).

− <S2>BS. These plots turn out to be surprisingly linear, and the best straight lines are shown
in Fig. 4.5(a) and Fig. 4.5(b) respectively. The extrapolated values of J, for Δ<S2> =1, are
0.16 cm−1 for 12 and 0.67 cm−1 for 13. From experiment, Wautelet et al20 concluded that the J
values for 12 and 13 are larger than the hfcc but extremely small in magnitude, less than 1 K.
The J values extrapolated here are in agreement with the experimental observations.
In principle, the magnetic exchange coupling constant J can also be determined by the
CASSCF methodology by using J = ES − ET. In practice, however, an explicitly detailed
CASSCF calculation can be performed only on very small species. For larger species such as
the diradicals 11−13, one can carry out a CASSCF calculation with only a handful of active

69
Chapter 4 Nitronyl Nitroxide

electrons in a handful of active orbitals. This limitation invariably results in a large value of
the calculated energy difference between the spin states, and the J values cannot be accurately

Table 4.10. Total energy from UB3LYP single point calculations on NN-pm-NN (13) in both
broken-symmetry (BS) and triplet (T) states using different basis sets.

Basis EBS(a.u.) ET(a.u.) EBS−ET (Δ<S2> −1) J (cm–1)a


<S2> <S2> (cal mol-1) x104

6-31G(d) −1605.8625392 −1605.8625728 21.08 37.26 7.40


1.078432 2.082158
6-31G(d,p) −1605.9099065 −1605.9099236 10.73 20.96 3.76
1.082187 2.084283
6-31+G(d,p) −1605.9613842 −1605.9613969 7.97 9.98 2.8
1.075813 2.076811
6-311G(d) −1606.2261714 −1606.2262023 19.39 34.18 6.79
1.077158 2.080576
Extrapolatedb 0.67

a
From Eq. (2.4). b Fig. 4.5 (b).

determined. In fact, Table 4.11 shows the results from the CASSCF calculations on
diradicals 11−13 by using 10 active electrons in 10 active orbitals. The spin state energy
difference (ET−ES) comes out to be very large in every case, of the order of a few kcal mol−1.
Nevertheless, the CASSCF results definitely identify the ground state spin. Table 4.11 clearly
shows IN-2p-IN (11) to have a singlet ground state whereas IN-pm-IN (12) and NN-pm-NN
(13) have triplet ground states.

4.6. Conclusions
A series of bis-nitrotronyl nitroxide diradicals with ten different conjugated couplers
have been investigated by broken-symmetry density functional treatment. The computed
magnetic exchange coupling constants are in very good agreement with the reported values.
Moreover, J values for 3, 4, 8 and 9 are predicted here (−230 cm-1, −136 cm-1, −148 cm-1 and
−161 cm-1, respectively). Sometimes it becomes necessary to explicitly compute the overlap
integral between the two magnetically active orbitals to calculate the exchange coupling
constant accurately by the broken-symmetry approach. The α-HOMO and β-HOMO in the
BS state are generally found to be magnetic orbitals.

70
Chapter 4 Nitronyl Nitroxide

3 8

6
J (cm )

2
−1

J (cm )
−1
4

1
2

0 0
0 2 4 6 8 10 0 10 20 30 40
2 4 4
(Δ<S >−1)x10 2
(Δ<S >−1)x10

4.5(a) 4.5.(b)

Figure 4.5. The best straight line plot of the computed J values against (Δ<S2>−1). The final J is
obtained by extrapolating the straight line to Δ<S2> = 1. We get (a) IN-pm-IN, J= 0.16
cm−1, and (b) NN-pm-NN, J=0.67 cm−1. The standard deviations are (a) 0.25 and (b)
0.52 in cm−1.

Table 4.11. Results from CASSCF (10,10) calculations with different basis sets for the diradicals
11−13.

Molecule Basis set Energy (a.u.) ES −ET J (cm−1)


(kcal mol−1)
Single (S) Triplet (T)

IN-2p-IN 6-31G −1445.7448 −1445.7237 −13.24 −2323


(11) 6-311G(d) –1446.6345 −1446.6331 −0.88 −154.1

IN-pm-IN 6-31G −1445.7012 −1445.7788 48.69 8542


(12) 6-311G(d) –1446.5921 –1446.6342 26.42 4632

NN-pm-NN 6-31G −1595.2477 −1595.2859 23.97 4205


(13) 6-311G(d) Convergence failure

In conjugated systems, the magnetic interaction is mainly transmitted through the π-


electron conjugation. The strength of antiferromagnetic interaction decreases with the increase
in the length of conjugated couplers. Conjugated linear couplers are more efficient
antiferromagnetic couplers than the aromatic ones of similar length. As the aromaticity of the
spacer decreases, the magnitude of the antiferromagnetic coupling constants increases. In
general, aromaticity favors the ferromagnetic trend. The diradicals with m-couplers are
undoubtedly ferromagnetic. The shape of the SOMOs as well as the rule of spin alternation in

71
Chapter 4 Nitronyl Nitroxide

the UHF emerge as two robust guidelines for the prediction of the qualitative nature of the
intramolecular magnetic interaction in bis-nitronyl nitroxide diradicals.
The intra-molecular magnetic exchange interaction between the radical units is
antiferromagnetic with J = −3.60 cm−1 for the diradical 11 and ferromagnetic for the diradical
12 and 13 with coupling constants 0.16 cm−1 and 0.67 cm−1 respectively. These calculated
values are in excellent agreement with the experimental results. The calculated J is also
dependent on the basis set so that it requires higher basis sets to get reliable results, especially
when the interaction is very weak.

References
1. Alies, F.; Luneau, D.; Laugier, J.; Rey, P. J. Phys. Chem. 1993, 97, 2922.
2. Schaftenaar, G.; Noordik, J. H. "Molden: a pre- and post-processing program for molecular and
electronic structures", J. Comput.-Aided Mol. Design, 2000, 14, 123.
3. Flükiger, P.; Lüthi, H. P.; Portann, S.; Weber, J. MOLEKEL, v.4.3; Scientific Computing: Manno,
Switzerland, 2002-2002. Portman, S.; Lüthi, H. P. CHIMIA 2000, 54, 766.
4. Barone, V. Recent Advances in Density Functional Methods; Part I; Ed. D. P. Chong; World Scientific
Publ. Co., Singapore, 1996.
5. (a) Ziessel, R.; Stroh, C.; Heise, H.; Köehler, F. K.; Turek, P.; Claiser, N.; Souhassou, M.; Lecomte, C.
J. Am. Chem. Soc. 2004, 126, 12604. (b) Stroh, C.; Ziessel, R.; Raudaschl-Sieber, G.; Köehler, F.;
Turek, P. J. Mater. Chem. 2005, 15, 850.
6. Ali, Md. E.; Vyas, S.; Datta, S. N. J. Phys. Chem. A 2005, 109, 6272.
7. (a) Trindle, C.; Datta, S. N. Int. J. Quantum Chem. 1996, 57, 781. (b)Trindle, C.; Datta, S. N.; Mallik, B.
J. Am. Chem. Soc. 1997, 119, 12947.
8. Caneschi, A.; Chiesi, P.; David, L.; Ferraro, F.; Gatteschi, D.; Sessoli, R. Inorg. Chem. 1993, 32, 1445.
9. Ziessel, R.; Ulrich, G.; Lawson, R. C.; Echegoyen, L.; J. Mater. Chem. 1999, 9, 1435.
10. Shiomi, D.; Tamura, M.; Sawa, H.; Kato, K.; Kinoshita, H. Syn. Metals 1993, 56, 3279.
11. Mitsumori, T.; Inoue, K.; Koga, N.; Iwamura, H.; J. Am. Chem. Soc. 1995, 117, 2467.
12. Hoffmann, R.; Zeiss, G. D.; Van Dine, G. W. J. Am. Chem. Soc. 1968, 90, 1485.
13. Constantinides, C. P.; Koutentis, P. A.; Schatz, J. J. Am. Chem. Soc. 2004, 126, 16232.
14. Zhang, G.; Li, S.; Jiang, Y. J. Phys. Chem. A 2003, 107, 5373.
15. Hay, P. J.; Thibeault, C. J.; Hoffmann, R. J. Am. Chem. Soc. 1975, 97, 4884.
16. (a) D’Anna, J. A. ; Wharton, J. H. J. Chem. Phys. 1970, 53, 4047. (b) Jurgens, O. ; Cirujeda, J. ; Mas,
M. ; Mata, I. ; Cabrero, A. ; Vidal-Gancedo, J. ; Rovira, C.; Molins, E. ; Veciana, J. J. Mater. Chem.
1997, 7, 1723 . (c) Zeissel, R. ; Ulrich, G. ; Lawson, R. C. ; Echegoyen, L. J. Mater. Chem. 1999, 9,
1435 . (d) Shiomi, D. ; Sato, K. ; Takui, T. ; Itoh, K. ; Tamura, M. ; Nishio, Y. ; Kajita, K. ; Nakagawa,
M. ; Ishida, T. ; Nogami, T. Mol. Cryst. Liq. Cryst. 1999, 335, 359.

72
Chapter 4 Nitronyl Nitroxide

17. (a) Luckhurst, G. R. In Spin Labeling. Theory and applications ; Berliner, J. L., Ed. ; Academic Press:
New York, 1976; p 133 ff. (b) Luckhurst, G. R. ; Pedulli, G. F. J Am. Chem. Soc. 1970, 92, 4738 ; (c)
Dulog, L. ; Kim, J. S. Makromol. Chemie 1989, 190, 2609.
18. Cirujeda, J.; Vidal-Gancedo, J.; Jürgens, O.; Mota, F.; Novoa, J. J. ; Rovira, C.; Veciana, J. J. Am. Chem.
Soc. 2000, 122, 11393.
19. Ali, Md. E.; Datta, S. N. J. Phys. Chem. A 2006, 110, 2776.
20. Wautelet, P.; Le Moigne, J.; Videva, V.; Turek, P. J. Org. Chem. 2003, 68, 8025.

73
Chapter 5
Influence of Aromaticity in Intramolecular Magnetic
Coupling
In this Chapter we discuss the prediction of the intramolecular magnetic exchange coupling
constant (J) for nitronyl nitroxide (NN) diradicals with different linear and angular polyacene
couplers. For the linear acene couplers, J initially decreases with increase in the number of
fused rings. But from anthracene coupler onwards, the J value increases with the number of
benzenoid rings due to an increasing diradical character of the coupler moiety. The J value
for the diradical with a fused bent coupler is always found to be smaller than that for a
diradical with a linear coupler of the same size. Nuclear independent chemical shift (NICS) is
calculated and it is observed that the average of the NICS values per benzenoid ring in the
diradical is less than that in the normal polyacene molecule. An empirical formula for the
magnetic exchange coupling constant of a NN diradical with an aromatic spacer is obtained
by combining the Wiberg bond order (BO), the angle of twist (φ) of the monoradical (NN)
plane from the plane of the coupler, and the NICS values. A comparison of the formula with
the computed values reveals that from tetracene onwards, the diradical nature of the linear
acene couplers becomes prominent thereby leading to an increase in the ferromagnetic
coupling constant. Isotropic hyperfine coupling constants are calculated by using polarized
continuum model for the diradicals in different solvents and in vacuum. In the last section of
this Chapter we have discussed the substituent effects on the m-phenylene couplers and the
verification of the empirical relationship is also described.
Chapter 5 Influence of Aromaticity

5.1. Introduction
Nitronyl nitroxide (NN) radicals have become the natural choice in molecular
magnetism since they are stable at ordinary conditions of temperature and pressure and also
have cooperative magnetic properties.1 These radicals are well characterized from structural
and spectroscopic viewpoints. The strong localization of the unpaired electron of NO makes
the NN radicals as ideal ferromagnetic precursors.2 In Chapter 4, we have noticed that the
π-conjugated linear spacers are, as couplers, stronger than the aromatic ones. We have also
noticed that the aromaticity of the coupler plays a major role in controlling the strength of
magnetic interaction. The m-phenylene species is known to be one of the best ferromagnetic
couplers. In the present work, we investigate the magnetic properties of eleven
ferromagnetically coupled NN diradicals with linear and angular polyacene couplers (Figure
5.1). The polyacenes are aromatic hydrocarbons with benzenoid rings. They have been
extensively investigated for their electronic properties, molecular structure, and aromaticity.
Pentacene has attracted a special attention as an active organic semiconductor molecule.3 The
larger polyacenes are predicted to be conductors with nearly zero band gap.4 The objective
of this work is to investigate the intramolecular ferromagnetic interaction mediated by
polyacene spacers.
In the latter part of this chapter we investigate 27 substituted m-phenylene based
nitronyl nitroxide diradicals (Figure 5.2). The object of this portion is finding out the effect of
subtitutent of m-phenylene in intramolecular ferromagnetic coupling.

5.2. Computational Methodology


Molecular geometries are optimized at the ROHF level using 6-311G(d,p) basis sets.
The magnetic exchange coupling constants are calculated using the spin-polarized
unrestricted DFT methodology. The magnetic exchange coupling constants are calculated by
using UB3LYP/6-311+G(d,p) methodology and following equations Eq. (2.4) and Eq. (2.6).
To study the effect of aromaticity of the coupler on the magnetic exchange interaction,
the nucleus-independent chemical shift (NICS) are calculated by B3LYP/GIAO methodology
for all the aromatic rings in each diradical. The NICS values are calculated at the centre of
the rings [NICS(0)]. But the σ framework of C−C and C−H affects the π electrons and hence

74
Chapter 5 Influence of Aromaticity

NICS is also calculated at a point 1 Å above the ring center [NICS(1)] where the π-electron
density is known to be maximum.

1 6 1Ο
Ο Ο 6
Ο
2 N A 7 + A B +
3 8 N 2N 3 8 7N

+ 9 + 9
N N N 10 N
4 Ο Ο 4 Ο Ο
5 10 5 2
1

1
1
Ο 6
Ο Ο 6Ο
A B C 7 + 2N A B C D 8 7 N+
2N 8 N 3
3

+ 9 + 9 N
N 10 N N
4 5Ο
10 Ο 4 Ο 10 Ο
3 5 5 4
A
C
1
Ο
2
1
Ο 6Ο
N B
3 C
A B C D E A 6
2N 3 8 7 N+ 4 + Ο
N 8 7 +

N
+ 9 9
N N 10
4 Ο Ο
N
5 10 Ο 6
5

1
Ο
B C 2
1Ο Ο6 N B
7 3 A C 6Ο
A D +
2N 8 N
3 + 8 7 +
N D N
9 4
+ N Ο 9
N 10 Ο 5 N
10
4 5Ο 8 Ο
7

B B C
C
1Ο 6Ο 1Ο 6Ο
A A D
D + 3 8 7N
+
2N 3 8 7N 2N

H H 9
+ +
9N N N
N
Ο 10 4 Ο 10
Ο
4
5 9 Ο 5
10

B C
6
1Ο Ο
A D 7
2 8
+
N 3 N
E 9
+ F N
N 10
4 Ο Ο
5

11
Figure 5.1. Diradicals under investigation. The benzene rings are identified by alphabets A, B, C,
etc. for the shake of convenience of discussion.

75
Chapter 5 Influence of Aromaticity

The isotropic hyperfine coupling constant (hfcc), which is essential in experiment to


characterize the radical systems and to predict the intramolecular exchange interaction, is also
calculated at B3LYP/EPR-II/C-PCM level. The hfcc values are first determined for the
diradical in vacuum. EPR parameters are strongly solvent dependent. To account the solvent
effect, hfcc’s are also calculated using conductor-like polarizable continuum model (CPCM).
Three solvents have been considered. These are the nonpolar solvent benzene (ε=2.25), the
moderately polar and aprotic solvent acetonitrile (ε=36.64), and the polar and protic solvent
water (ε=78.39).

5.3. Results and Discussion


Table 5.1 shows the calculation of J from Eqs. (2.4) and (2.6). The decreasing order
of the J from 1 to 3 is in agreement with our general observation that magnetic exchange
interaction in NN diradical with linear conjugated couplers decreases with the increase in the
length of the coupler.5 The reason for the deviation of 3-5 is that the larger oligoacenes
possess open-shell singlet ground states,6a that is, these acenes are diradicals with disjoint
nature.
The J value decreases remarkably for the bent couplers. The coupler in 6 has 3 fused
rings. Those in 7-9 have 4 fused rings. The coupler in 10 has 4 fused aromatic and one fused
non-aromatic rings, while that in 11 has 7 fused aromatic rings. One consequence of being
bent is that the coupler fragments lose the disjoint diradical character. Also, they become
stronger aromatics as discussed later. Diradicals 3 and 6 are similar, but J is much smaller for
phenanthrene coupler (6) than the anthracene one (3). The J value further decreases for the
1,8 and 1,7 substituted pyrene couplers (7 and 8). The couplers of diradicals 2 and 7 are
similar but the value of J for 2 is more than twice that for 7. Again, 6 and 8 are similar in
length except for an additional ring in 8, but J is larger in 6 than in 8. The same trend, that is,
the decrease of the J value with the increase of conjugation in the bent aromatic coupler is
observed in the case of 9-11. Nevertheless, conjugation within the coupler is not the only
factor that determines the strength of the intramolecular exchange interaction.

76
Chapter 5 Influence of Aromaticity

5.3.1. Bond Order and Dihedral Angles


Wiberg bond index (order)7 is calculated by natural bond orbital (NBO) analysis
(implemented in Gaussian 03) at B3LYP/6-311+G(d,p) level. The calculated bond orders are
given in Table 5.2 along with the angle of rotation of the NN plane from the coupler plane (φ).

Table 5.1. Results from single-point broken-symmetry calculations at UB3LYP/6-311+G(d,p) levels


and the calculated J values. The triplet geometry is optimized. Legends: J(2) for GND
equation (2.4) with Smax=1, and J(4) from Yamaguchi equation (2.6).

Energy (a.u.) J(cm−1)


Diradical
(<S2>)
BS T Eq. (2.4) Eq. (2.6)
−1299.0066208 −1299.00 6716
1a 20.89 20.78
1.0703 2.0756
−1452.68288294 −1452.68295337
2 15.46 15.39
1.0706 2.0754
−1606.35279897 −1606.35286055
3 13.52 13.44
1.0720 2.0775
−1760.02026569 −1760.02032912
4 13.92 13.81
1.0742 2.0821
−1913.68945590 −1913.68953368
5 17.07 16.73
1.0767 2.0910
−1606.32389842 −1606.32393844
6 8.78 8.76
1.0946 2.0974
−1682.60289023 −1682.60291783 6.05
7 6.06
1.0706 2.0726
−1682.60180740 −1682.60183056
8 5.08 5.07
1.0683 2.0701
−1760.03222796 −1760.03224969
9 4.77 4.76
1.0825 2.0842
−1798.16513473 −1798.16515225
10 3.85 3.84
1.0707 2.0722
−1988.76716320 −1988.76716712
11 0.86 0.86
1.0660 2.0663

a
These values are reported in reference Ref. 5, and the observed J value is 20 cm−1.

The average bond order (BO) for the linear acene couplers (1-5) increases with the
increase of the number of phenyl rings in the coupler. A larger bond order generally favors a
greater conjugation with the radical centers, and hence a larger magnetic exchange coupling

77
Chapter 5 Influence of Aromaticity

constant. The rotation of the NN plane from the plane of the coupler (φ) has an opposite
effect, that is, if φ increases, J decreases because of the lesser conjugation. In 1-3, J decreases
although BO increases and φ decreases whereas for 3-5, J increases with the size of the
coupler along with the increase in φ.

Table 5.2. The calculated Wiberg bond order at B3LYP/6-311+G(d,p) level for the NN–coupler bond
and the average dihedral angle between the NN and coupler planes.

Diradical Bond Ordera Dihedral Angleb


r1 r2 Average φ1 φ2 Average

1 1.05 1.05 1.05 34.74 32.33 33.54


2 1.07 1.07 1.07 25.76 25.78 25.77
3 1.08 1.07 1.08 26.45 22.87 24.66
4 1.07 1.08 1.08 26.87 22.71 24.79
5 1.07 1.08 1.08 27.06 22.61 24.84

6 1.07 1.05 1.06 5.72 43.65 24.68


7 1.07 1.02 1.05 23.01 54.22 38.61
8 1.07 1.03 1.09 22.90 53.87 38.38
9 1.07 1.07 1.07 24.22 23.16 23.69
10 1.07 1.07 1.07 23.96 25.22 24.59
11 1.02 1.02 1.02 55.32 55.31 55.31

a
r1 and r2 are bond lengths between the benzenoid ring and the two NN radicals.
b
φ1 and φ2 are the angles of twist of the two NN moieties from the plane of the coupler.

The planes of the two NN moieties are asymmetrically twisted for the angular
diradicals 6-8. One of the NN planes undergoes a large twist, and this fact is also reflected in
BO. The BO and φ are consistent with the trend in calculated J values for the diradicals 6-8,
and a similar trend is observed for 9-11. For the highly planar and conjugated coupler
coronene (11), the very less J value (0.86 cm−1) is due to the extremely large angle of twist
(φ=55.31°), basically a stereo-electronic effect.

5.3.2. Nuclear Independent Chemical Shift (NICS)


NICS(0) and NICS(1) are calculated at GIAO-B3LYP/6-311+G(d,p) level for
different six membered rings in each coupler. NICS is an accepted measure of aromaticity.
The benzenoid rings are denoted as A, B, C, etc. in Figure 5.1 and the corresponding values of
NICS(0) and NICS(1) are given in Table 5.3.

78
Chapter 5 Influence of Aromaticity

Table 5.3. The calculated NICS values at the center of the aromatic rings for diradicals 1-11.

Diradical NICS A B C D

NICS(0) −7.51

1 NICS(1) −9.24

NICS(0) −7.60 −7.57

2 NICS(1) −9.62 −9.68

NICS(0) −6.47 −10.3 −6.28

3 NICS(1) −8.61 −12.22 −8.41

NICS(0) −5.15 −10.5 −10.38 −5.04

4 NICS(1) −7.51 −12.4 −12.29 −7.35

NICS(0) −3.87 −9.60 −11.40 −9.61

5a NICS(1) −6.62 −11.62 − 13.64 −

11.51

NICS(0) −7.95 −4.66 −8.06

6 NICS(1) −9.96 −7.72 −10.16

79
Chapter 5 Influence of Aromaticity

NICS(0) −10.41 −2.86 −10.85 −2.72

7 NICS(1) −12.19 −6.29 −12.58 −5.95

NICS(0) −10.68 −2.82 −10.32 −2.93

8 NICS(1) −12.42 −6.38 −12.13 −6.33

NICS(0) −8.04 −5.63 −5.61 −8.20

9 NICS(1) −8.85 −8.06 −8.81 −11.12

NICS(0) −7.36 −5.21 −5.17 −7.49

10 NICS(1) −9.37 −8.06 −8.07 −9.48

NICS(0) − 8.65 − 8.37 − 8.50 −8.95

11b NICS(1) −10.87 −10.85 −10.92 −

10.67

a
NICS(0) and NICS(1) for ring E in species 5 are −3.87 and −6.34 respectively; b NICS(0) and NICS(1) for ring
E in species 11 are −8.38 and −10.83 respectively and for ring F are −8.54 and −10.93.

The linear polyacene molecules have already been investigated by Schleyer et al.8
using the same methodology and basis sets. These authors observed that the terminal rings
have less benzenoid character as the size of the linear acene increases. For the angular
acenes, however, the central rings have a reduced benzenoid character except in the

80
Chapter 5 Influence of Aromaticity

thoroughly aromatic molecule 11. These trends are exactly preserved for the acene couplers
in the diradicals under investigation (Table 5.3).
Table 5.4 shows that the average NICS(1) for a coupler is always less than that for the
normal acene molecule. The difference between NICS(1) of a NN diradical and that of the
corresponding acene molecule is written as ΔNICS. The loss of aromaticity in the coupler
moiety is due to the participation of the conjugated π-electrons in the magnetic exchange
phenomenon. We notice that J is proportional to the fractional change of NICS(1) from the
parent acene, that is, ΔNICS : |NICS(1)| for acene. It is also generally proportional to the
Wiberg bond order BO, cos φ1 and cos φ2. As we discussed in Chapter 4 for linear aliphatic
couplers, the absolute magnitude of atomic spin density approximately varies as 1/(N+1)
where N is the number of conjugated atoms in the coupler, and J is approximately
proportional to 1/(N+1). Similarly, here, J will be further proportional to a factor of 1/(n+1)
where n is the number of benzenoid rings in the ployacene coupler. These proportionalities
can be coupled together to write the qualitative expression

( ΔNICS )( BO ) cosφ1 cosφ 2


J = A× (5.1)
( n + 1)( NICS)

where NICS in the denominator is the absolute magnitude of NICS(1) for the parent acene.
The proportionality constant A is found by considering the experimental value J = 20 cm−1 for
the m-phenylene coupler (with n=1).9 We get A = 426.5 cm−1. The J values calculated from
Eq. (5.1) are given in Table 5.4. It is seen that Eq. (5.1) produces a rough estimate of J, but
for the linear polyacenes, the estimate grows progressively worse from 3 to 5.
For the liner acenes, the average NICS(1) per benzenoid ring increases with the size of
the coupler (Table 5.4). It is also evident that the diradical character increases with the
coupler size.6a Introducing the effective value (1-χd)NICS in place of NICS in the
denominator of Eq. (5.1) for the diradicals with linear couplers, and using the scaling of the
calculated J values by the multiplicative factor 0.9625 (=20.0/20.78), we find the deviation
parameter χd as 0.0, 0.12, 0.17, 0.38 and 0.56 respectively for n equal to 1–5. The deviation
parameter reflects the trend of the increasing diradical character.
81
Chapter 5 Influence of Aromaticity

Table 5.4. The calculated NICS(1) values for the diradicals and the corresponding acene
molecules, ΔNICS(1) and the J value calculated from Eq. (6) and estimated from Eq.
(5.1).

Diradical Average NICS(1) ΔNICS(1) JGND J

Couplersa Acenesb Eq. Eq.

(2.5 (5.1

) )

1 −9.24 −10.60c 1.36 20.78 20.00

2 −9.56 −10.80c 1.15 15.39 13.10

3 −9.75 −11.00c 1.25 13.44 10.72

4 −9.89 −11.10c 1.21 13.81 8.22

5 −9.94 −11.20c 1.25 16.73 7.01

6 −9.28 −9.94c 0.66 8.76 5.42

7 −9.25 −10.62d 1.37 6.05 6.19

8 −9.32 −10.62d 1.30 5.07 5.95

9 −9.21 −9.59d 0.38 4.76 3.02

10 −8.75 −9.27d 0.52 3.84 4.23

11 −10.83 −12.17d 1.34 0.86 2.22

a
In NN diradical; bAcene molecule without any NN radical as substituent; cSchleyer et al. Ref. 8; dCalculated at
GIAO-B3LYP/6-311+G(d,p) level, our work.

82
Chapter 5 Influence of Aromaticity

The deviation cannot be straight-forwardly applied to the bent couplers where the
central rings are less aromatic. Also, the variation of NICS(1) is not smooth like that in linear
couplers because of the zwitterionic contributions in bent acenes.6b The J values estimated
by Eq. (5.1) in this case are generally in better agreement with the calculated values.

5.3.3. SOMO-SOMO Energy Splitting


The energies of the SOMOs are calculated at UB3LYP/6-311+G(d,p) level. The
(ε1SOMO−ε2SOMO) energies are very low except for 6. The difference (ε1SOMO−ε2SOMO)
decreases with increase in length of the linear acenes (1-5). The degeneracy of the SOMOs
for 2, 10 and 11 arises accidentally. The molecular point group is C2 (see Table 5.5) that is
abelian. The rest of the diradicals undergo a distortion from this symmetry. That all the
diradicals have SOMO-SOMO energy difference less than 1.5 eV and all have ferromagnetic
ground states are in agreement with the empirical rule proposed by Hoffman.

5.3.4. Isotropic Hyperfine Coupling Constant (hfcc)


The polarized continuum model (PCM) has been successfully applied to the
investigation of isotropic hyperfine coupling constant (aN) of organic radicals in solution.
The solute-solvent interaction can change aN values by modifying the local spin density. In
this work, we have calculated aN values for all four equivalent N-atoms in each diradical.

Table 5.5. The SOMO-SOMO energy splitting at UB3LYP/6-311+G(d,p) level for the triplet state.

Diradical E(SOMO1) E(SOMO2) ΔE(SOMO)


(a.u.) (a.u.) (a.u.)

1 −0.19697 −0.19183 0.0051


2 −0.19836 −0.19831 0.0001
3 −0.19958 −0.19806 0.0015
4 −0.19920 −0.19421 0.0050
5 −0.19895 −0.18585 0.0131
6 −0.22288 −0.16391 0.0590
7 −0.19963 −0.18928 0.0104
8 −0.19970 −0.19223 0.0075
9 −0.19819 −0.18973 0.0085
10 −0.19837 −0.19778 0.0006
11 −0.19359 −0.19358 0.0000

83
Chapter 5 Influence of Aromaticity

The experimental values lie in the range of 3-4.5 G for diradicals with different
couplers.10 In this work, we have calculated the hfcc for the diradical in gas phase as well as
in three different solvents. The calculated average hfcc for the N-atoms are given in Table
5.6. The average aN values for the nitrogen atoms are in good agreement with the
experimental values for general nitronyl nitroxide diradicals. The calculated values indicate
that there is a preference for the spin density to localize on one of the N-atoms in each NN
moiety.
Solvent plays an important role in hfcc. In all the species, the hfcc values for N-atoms
increase with the increase in dielectric constant. For linear acenes (1-5), the average hfcc
value increases as the coupler size increase. However, a straightforward correlation of the aN
values with the calculated J remains missing.

5.4. The m-Phenylene Couplers


In this section we will further check the validity of Eq. (5.1) and the substituent effects
on phenyl ring of the m-phenylene coupler, which is know as the best ferromagnetic spacer.
Here we have investigated the magnetic properties of 27 substituted m-phenylene based
Nitronyl Nitroxide diradicals. We have considered nine different constituents on the m-
phenylene ring, namely, −COOH, −F, −Cl, −NO2, −Br, −OH, −NH2, −Ph and −CH3, in order
of decreasing −I effect (electron withdrawing power). Each substituent occupies three unique
positions on the ring (Figure 5.2). The role of aromaticity of the coupler unit is studied
explicitly. All molecules are new, with the exception of 15b which has been recently
synthesized.11

Table 5.6. Calculated average isotropic hyperfine coupling constant (hfcc) for nitrogen atoms of the
diradical in different environments.

Diradicals Gas phase Benzene Acetonitrile Water

1 2.78 2.96 2.95 2.97

2 2.90 3.15 3.23 3.25

3 3.26 3.25 3.37 3.35

4 3.26 3.23 3.38 3.40


84
Chapter 5 Influence of Aromaticity

5 3.07 3.29 3.37 3.40

6 2.76 3.15 3.23 3.38

7 2.82 2.90 2.95 3.04

8 2.96 3.02 3.12 3.14

9 2.78 2.80 2.94 2.95

10 3.09 3.19 3.33 3.29

11 2.79 2.86 2.92 2.94

5.4.1. Calculations
The calculated JGND and JY for all the species are given in Table 5.7. The coupling
constant is found to follow the order a < b < c in every case. The maximum variation in J is
found for the species with substitution at the common ortho position for both NN fragments (a
isomers). The calculated J is in the range of 3.59−8.34 cm−1, except for 15a, 16a and 19a. In
the last three cases, the ground state is a singlet, and the magnetic interaction is
antiferromagnetic in nature. For the ortho-para (b) species, J varies from 7.07 cm−1 to 14.87
cm−1 except for 17b and 19b. For 17b, J is 20.78 cm−1 and for 19b, it is 1.95 cm-1. The meta
substitution (c) always yields a J value in the range of 20.11-24.05 cm−1, except for 16c where
it is 27.11 cm−1.
The decreasing trend of −I effect reveals no clear-cut systematics for any substituent
position, as electron delocalization is the major factor for determining the coupling constant.
For the single atom substituents F, Cl and Br (species 13, 14 and 16), however, a pattern
clearly emerges. When the substituents are in ortho positions (a and b), the coupling constant
decreases with the decreasing −I effect, whereas for meta position (c), J clearly increases.
This can be understood from spin alternation. An increase in electron density at the ortho
positions leads to a destructive interference. This explanation would fail in the case of a
substituent with more than one atom, as the bond(s) of the substituent would be involved in
delocalization, hydrogen bonding, steric effect, etc.

85
Chapter 5 Influence of Aromaticity

The angles of rotation (φ1 and φ2 as in Figure 5.3) of NN from the plane of the m-
phenylene coupler are given in Table 5.8 which clearly indicates that J decreases as φ
increases. This is because the conjugation between the NN and coupler fragment decreases
with the increase in φ,12 and is in agreement with the work in previous Chapter. The angle φ1
and φ2 are constant in almost every case except 16c and 17c for the substitution at the
common meta position (c) and vary between 26°-33°. The larger values for 1c are due to the
steric effect of the bulkier –COOH group. The reason for the very small and unsymmetrical φ
values for 16c and 17c remains unclear. The larger dihedral angles for a and b isomers are
mainly due to stereo-electronic effects. We find that the calculated J is approximately
proportional to cosφ1 cosφ2.
Wiberg bond order (BO) for NN-Ph for all the cases is greater than 1.0 except for 15a
and 18a. The larger BO (>1.0) indicates a larger double bond character and hence a greater
conjugation, thus a larger J. It is fair to estimate that J varies linerly with BO. The BO for
15a and 18a are discussed later. An estimate of J can be obtained from Eq. (5.1). This
equation is valid only if steric and other constraints are absent. We have recently shown that

86
Chapter 5 Influence of Aromaticity

CO2H

O O
CO2H
O O
O O
N N N N
N N
N CO2H N N N
O O N N O O
O O
12a 12c
12b

O O F
O O O O
N N N N
N N
N F N N N
O O N N O O
O O
13a 13b 13c

Cl
Cl
O O O O
O O
N N N N
N N
N Cl N N N
O O N N O O
O O
14a 14c
14b
NO2
NO2
O O O O
O O
N N N N
N N

N NO2 N N N
O O N N O O
O O

15a 15b 15c


Br
Br
O O O O
O O
N N N N
N N
N Br N N N
O O N N O O
O O
16a 16b 16c

OH

O O OH O O
O O
N N N N
N N

N OH N
N N N N
O O O O
O O

17a 17c
17b

87
Chapter 5 Influence of Aromaticity

NH2

O O NH2 O O
O O
N N N N
N N

N NH2 N N N
O O N N O O
O O

18a 18b 18c

Ph
Ph
O O O O
O O
N N N N
N N

N Ph N N N
O O N N O O
O O

19a 19b 19c

CH3
CH3
O O O O
O O
N N N N
N N

N CH3 N N N
O N N O O
O O O

20a 20b 20c

Figure 5.2. The diradicals under investigation.

O 11 5 16 O
6 4
φ1 φ2 +
N10 1 3 N 15
2
9 14
+ 13 N
N 8
O O
7 12

Figure 5.3. Schematic representation of m-phenylene-coupled nitronyl nitroxide diradical.

this equation gives an estimate of J for polyacenes.12 For the present systems, we find A to
be about 426.5 cm−1. The calculated coupling constant (Jcal) using equation (5.1) matches
well with calculated values of J for the c group species, except for 18c and 19c where the
calculated ΔNICS(1) is too small and large respectively. For the ortho species, steric
constraint, hydrogen bnding, etc. leads to prominent deviations. The J values estimated from
Eq. (5.1) are given in Table 5.8 along with all the parameters involved in the equation. The

88
Chapter 5 Influence of Aromaticity

above equation produces the absolute magnitude of J and does not determine the nature of the
magnetic exchange interaction.

Table 5.7. Calculated exchange coupling constant (J) at UB3LYP/6-311G(d,p) level.


Diradicals Energy(a.u.) JGND JY
<S2>
BS T (cm−1) (cm−1)

12a −1487.6248083 −1487.6248263 3.94 3.94


1.0587 2.0595
12b −1487.6306294 −1487.6306716 9.24 9.26
1.069612 2.072114
12c −1487.6362267 −1487.6363185 20.11 20.01
1.0675 2.0727

13a −1398.2643806 −1398.2644187 8.34 8.32


1.0698 2.0719
13b −1398.2711268 −1398.2711833 12.37 12.35
1.071 2.0737
13c −1398.272202 −1398.2723105 23.76 23.61
1.0676 2.0740

14a −1758.6141617 −1758.6141781 3.59 3.59


1.0701 2.0710
14b −1758.6220774 −1758.6221191 9.13 9.11
1.0683 2.0705
14c −1758.6273742 −1758.627484 24.05 23.89
1.0677 2.0743

15a −1503.45099658 −1503.45083192 −36.14 −35.35


1.0331 2.0516
15b −1503.55431367 −1503.55437103 12.59 12.55
1.0723 2.0758
15c −1503.56186522 −1503.56196921 22.82 22.68
1.0678 2.0740

16a1 −1311.5478844 −1311.5478825 −0.40 −0.40


1.0705 2.0709
16b1 −1311.55683448 −1311.55686668 7.07 7.06
1.0710 2.0727
16c1 −1311.55832363 −1311.55844713 27.11 26.91
1.0646 2.0719

17a −1374.2588476 −1374.2588652 3.85 3.85


1.0578 2.0587
17b −1374.26158508 −1374.26167978 20.78 20.68
1.0727 2.0777
17c −1374.2448564 −1374.2449509 20.70 20.58
1.0559 2.0616

18a −1354.3834627 −1354.3835035 8.93 8.37


1.0634 2.0656

89
Chapter 5 Influence of Aromaticity

18b −1354.3929839 −1354.3930518 14.87 14.81


1.0624 2.0661
18c −1354.3814049 −1354.3815049 21.9 21.77
1.0663 2.0721

19a −1530.1020791 −1530.1020719 −1.58 −1.58


1.0639 2.0637
19b −1530.1105927 −1530.1106016 1.95 1.95
1.0664 2.0669
19c −1530.1158724 −1530.1159792 23.39 23.12
1.0668 2.0731

20a −1338.3213797 −1338.3214042 5.4 5.4


1.0658 2.0671
20b −1338.3316078 −1338.3316635 12.2 12.16
1.0649 2.0679
20c −1338.3336551 −1338.3337603 23.04 22.89
1.0663 2.0725
1
Using lanl2dz basis set for Br. The 6-311+G(d,p) basis set for Br yields a JGND value of −0.53 cm−1 for 16a,
7.05 cm−1 for 16b and 27.07cm−1 for 16c.

Table 5.8. Calculated NICS(1), bond order (B.O.), angle of twist (φ), and the J value estimated from
Eq. (5) (Jcal ) and calculated from the BS-DFT approach (JGND).

NICS(1) ΔNICS(1) BO φ1 φ2 Jcal JGND

Diradicalsa Moleculesb (cm-1) (cm-1)

12a −9.71 0.44 1.01 47.48 63.50 5.63 3.94


12b −9.52 −10.15 0.63 1.03 28.82 55.11 13.66 9.24
12c −9.42 0.73 1.04 32.35 32.96 22.61 20.11

13a −9.54 0.85 1.02 46.78 55.81 13.70 8.34


13b −9.63 −10.39 0.76 1.04 23.23 54.91 17.14 12.37
13c −9.68 0.71 1.06 25.62 25.62 25.12 23.76

14a −9.40 0.65 1.01 60.80 61.12 6.56 3.59


14b −9.39 −10.05 0.66 1.04 21.69 61.55 12.89 9.13
14c −9.35 0.70 1.05 25.84 25.84 25.27 24.05

15a −9.08 1.22 0.65 51.25 38.37 16.11 -36.14


15b −9.80 −10.30 0.50 1.06 15.94 54.64 12.21 12.59
15c −9.46 0.84 1.05 26.85 28.54 28.62 22.82

16a −9.32 0.56 1.01 60.23 60.08 6.05 -0.40


16b −9.32 −9.88 0.56 1.03 15.19 69.89 8.26 7.07
16c −9.08 0.80 1.06 28.33 15.53 31.05 27.11

90
Chapter 5 Influence of Aromaticity

17a −8.76 1.06 1.03 68.94 38.55 13.33 3.85


17b −8.66 −9.82 1.16 1.05 25.93 40.65 36.09 20.78
17c −9.23 0.59 1.06 2.33 35.39 22.13 20.70

18a −7.97 0.96 1.04 49.50 47.34 20.99 8.93


18b −7.89 −8.93 1.04 1.06 22.60 44.43 34.71 14.87
18c −8.69 0.24 1.05 28.56 27.04 9.42 21.90

19a −9.44 0.19 0.99 82.88 79.97 0.18 −1.58


19b −8.90 −9.63 0.73 1.02 22.40 95.08 -2.70 1.95
19c −8.74 0.89 1.05 26.34 26.34 33.24 23.39

20a −9.42 0.65 1.01 55.79 56.62 8.60 5.40


20b −9.62 −10.07 0.45 1.04 22.57 54.08 10.74 12.20
20c −9.42 0.73 1.04 32.35 32.96 22.79 20.11

a
This work; b Ref. 8.

Wiberg bond index for phenyl−NN is small for 15a, 16a and 19a. Each B.O. reported
in Table 5.8 is the average of two phenyl-NN bond orders. The low values are indicative of
the occupation of σ* orbitals by the non-bonding electrons. This favors the Heitler-London
spin pairing, hence the antiferromagnetic coupling, especially in 15a. The formation of a
unique O−O bond in 15a is an interesting phenomenon (Figure 5.4) that is also responsible for
the large antiferromagnetic interaction. Hydrogen-bonding produces mostly structural effects
such as a reduction of the twist angle φ as in 17a and 17b (Figure 5.4), and also a small
amount of spin delocalization effect. The bulky phenyl group leads to an increase in the twist
angle in its neighborhood, (both φ1 and φ2 in 19a and only φ2 in 19b). Thus 19a becomes
antiferromagnetically coupled and 19b is only faintly ferromagnetic. The explanation of the
antiferromagnetic coupling in 16a is a conundrum. It varies from a combination of different
factors, and can be best understood from a comparison of the different factors involved in Eq.
(5.1).

5.5. Conclusions
The magnetic exchange coupling constants are calculated for eleven diradicals by
broken-symmetry density functional method. The coupling constant J is found to decrease for
the linear acene couplers from one to three benzenoid rings, but it increases from three to five

91
Chapter 5 Influence of Aromaticity

benzenoid rings. The α-HOMO and β-HOMO are not the only magnetically active orbitals
for 3-5. This happens due to the increase of the diradical character of the acene couplers.
The diradical character is lost in the bent couplers. The NICS value at the central rings of the
linear acene is high, while the terminal rings lose

Figure 5.4. The unique O-O bond formation observed for 15a. No such bond is formed in 15b.
Intramolecular H-bonding is found in 17a and 17b. (Similar hydrogen bonds are also
observed for 12a, 18a and 18b).

the benzenoid character. The J value increases with BO, and decreases with the increase in
the angle of twist of the NN mono-radicals from the coupler plane. The qualitatively
proposed equation (5.1) can give a fair estimate of J. Reliable aN values are obtained for the
diradicals in solution.
The magnetic exchange coupling in 24 out 27 studied m-phenylene diradicals are
predicted as ferromagnetic at UB3LYP/6-311+G(d,p) level. Species 15a, 16a and 19a are

92
Chapter 5 Influence of Aromaticity

antiferromagnetically coupled. The substitution at the meta position of m-phenylene NN


diradicals has little steric and hydrogen bonding effects. In the case of a larger twist of the
planes of the spin sources from the plane of the coupler, the nature of the interaction changes
from ferromagnetic to antiferromagnetic. In these cases, the B.O. becomes small (even less
than 1.0) due to a partial population of the σ* orbital. The NICS value is found to decrease
from the corresponding mono-substituted phenyl derivatives. The coupling constant can be
estimated from Eq. (5.1) provided that complications arising from sterio-electronic repulsion,
hydrogen bonding, ring formation, etc., are absent. This is found to be generally valid for NN
diradicals with m-phenylene spacer that has substitutions at the common meta position.

5.6. References
(1) (a) Blundell, S. J.; Pratt, F. L. J. Phys.: Condens. Matter 2004, 16, R771. (b) Luneau, D.; Rey, P.
Coord. Chem. Rev. 2005, 249, 2591.
(2) Blundell, S.J.; Pratt, F.L. J. Phys.: Condens. Matter 2004, 16, R771.
(3) (a) Hegmann, F. A.; Tykwinski, R. R.; Lui, K. P. H.; Bullock, J. E.; Anthony, J. E. Phys. Rev. Lett. 2002,
89, 227403. (b) Meng, H.; Bendikov, M.; Mitchell, G.; Helgeson, R.; Wudl, F.; Bao, Z.; Siegrist, T.;
Kloc, C.; Chen, C.-H. Adv. Mater. 2003, 15, 1090.
(4) Raghu, C.; Patil, Y. A.; Ramasesha, S. Phys. Rev. B 2002, 65, 155204.
(5) Ali, Md. E.; Datta, S. N. J. Phys. Chem. A 2006, 110, 2776.
(6) (a) Bendikov, M.; Duong, H. M.; Starkey, K.; Houk, K. N.; Carter, E.A., and Wudl, F. J. Am. Chem. Soc.
2004, 126, 7416. (b) Constantinides, C.P.; Koutentis, P.A.; Schatz, J. J. Am. Chem. Soc., 2004, 126,
16232.
(7) Wiberg, K. Tetrahedron 1968, 24, 1083.
(8) Schleyer, P.v.R.; Maerker, C.; Dransfeld, A.; Jiao, H.; Hommes, N.J.R.v.E.
J. Am. Chem. Soc. 1996, 118, 6317 . (b) Chen, Z.; Wannere, C.S.; Corminboeuf, C.; Puchta, R.; and
Schleyer, P.v.R. Chem. Rev. 2005, 105, 3842. (c) Schleyer, P.v.R.; Manoharan, M.; Jiao, H.; Stahl, F.
Org. Lett. 2001, 3, 3643.
(9) Shiomi, D.; Tamura, M.; Sawa, H.; Kato, K.; Kinoshita, H. Syn. Metals 1993, 56, 3279.
(10) Hoffmann, R.; Zeiss, G. D.; Van Dine, G. W. J. Am. Chem. Soc. 1968, 90, 1485.
(11) (a) Luckhurst, G. R. In Spin Labeling. Theory and applications ; Berliner, J. L., Ed. ; Academic Press:
New York, 1976; p 133 ff. (b) Luckhurst, G. R. ; Pedulli, G. F. J Am. Chem. Soc. 1970, 92, 4738 ; (c)
Dulog, L. ; Kim, J. S. Makromol. Chemie 1989, 190, 2609. (d) Catala, L.; Le Moigne, J.; Kyritsakas, N.
Rey, P.; Novoa, J. J.; Turek, P. Chem. Eur. J 2001, 7, 2466.
(12) Ali, Md. E.; Datta, S. N. J. Phys. Chem. A 2006, 110, 13232.

93
Chapter 6
Photomagnetism

In this Chapter, we investigate another aspect of molecular magnetism. We predict the


photo-switching magnetic properties of four substituted dihydropyrenes from density
functional broken-symmetry calculations. The magnetic exchange coupling constants differ
up to 17 cm−1 between the photo switched isomers. The intramolecular exchange interactions
are ferromagnetic in nature. The calculated coupling constants are much larger than those
reported earlier for photomagnetic organic molecules. We also report the calculated values of
J for a few synthesized diradicals.
Chapter 6 Photomagnetism

6.1. Introduction
Photochromism is the reversible photon-induced transition of a chemical species
between two different forms having different absorption spectra. Photochromic materials
change their geometries and physical properties with irradiation. They are useful in potential
photoswitching. If a photoswitchable molecule is used as a spin coupler between two
magnetic units, the magnetism of the resulting species can change upon irradiation.1
Perfluorocyclopentene is one of the widely studied photochromic spin couplers.
Matsuda et al. have synthesized a large number of nitronyl nitroxide diradicals with
perfluorocyclopentene.1-4 In these diradicals, the intramolecular exchange interaction is very
weak, and the coupling constant J is of the order of the hyperfine coupling constant (hfcc).
The J value differs nearly 150-fold between the open- and closed-ring isomers. Its absolute
magnitude is generally found to be < 10−3 cm−1 for open ring isomers and ∼ 10−2 cm−1 for
closed ring isomers except for NN diradicals with 1,2-bis(2-methyl-1-benzothiophene-3-
yl)perfluorocyclopentene where J equals to −0.76 cm−1 and −4.03 cm−1 respectively.1b,1d,1g As
|J| is very small in the ground state, the photomagnetic properties of these molecules are not
expected to find a great usage. This has led to the investigation of photo-excited states of
diradicals. Teki et al. have investigated the magnetic properties of excited states of nitronyl
nitroxide diradicals with diphenylanthracene coupler.5 Huai et al. have also investigated
similar excited states by theoretical means.6

hν1, Δ

hν2

CPD DDP

Figure 6.1. Conversion of CPD to DDP.

The substituted pyrene molecule exists in two different forms, namely,


cyclophanediene (CPD) and dihydropyrene (DDP) as shown in Figure 6.1.7 The restricted life
time of CPD limits the utility of these molecules. The thermal return of CPD to DDP belongs
to the category of Woodward-Hoffman (W-H) orbital symmetry forbidden processes.

92
Chapter 6 Photomagnetism

R2 hν1, Δ R2
R1 R1
R2
hν2 R2
R1 R1
a b
O O
N+ N R1 R2
N N 1a, 1b -NN -CH3
O
2a, 2b -NN -CF3
-NN -IN 3a, 3b -IN -CH3
4a, 4b -IN -CF3

Figure 6.2. Diradicals under investigation.

Nevertheless, the barrier created from the correlation of the occupied reactant orbitals
with the virtual product orbitals and vice versa is not too high for CPD→DDP conversion.
Recently, Williams et al. have found that proper substitutions can increase the activation
barrier to hinder the thermal conversion.8 In this work, we have investigated the ground-state
photomagnetic properties of nitronyl nitroxide diradicals and imino-nitroxide diradicals with
substituted pyrene couplers. The four sets of diradicals are illustrated in Figure 6.2. One
novelty of this work is in the choice of the coupler. To our knowledge, these molecules have
not been synthesized so far. The J values of the isomers differ by 4.7-9.6 times for each pair.
The magnitude of J for the closed ring isomers is significantly large, which constitutes the
second novelty. Besides, the points of attachment of the NN and IN groups are decided from
the rule of spin alternation9 such that the resulting diradicals are ferromagnetic in nature.

6.2. Technical Details and Results


The theoretical evaluation of the magnetic exchange coupling constant have been
performed using broken-symmetry (BS) density functional (DFT) methodology proposed by
Noodleman. The J value is evaluated using the so called GND [Eq.(2.5)] and Yamaguchi
[Eq.(2.7)] equations.
The molecular geometries of all the eight species (1-4 a and b) are optimized at
ROHF/6-31G(d,p) level using Gaussian 03 software. The optimized molecular geometry for
1a and 1b are shown in Figure 6.3. The magnetic exchange coupling constants, which are

93
Chapter 6 Photomagnetism

calculated at UB3LYP/ 6-311+G(d,p) level, are given in Table 6.1 for all the species.

1a

1b

Figure 6.3. Optimized geometries for 1 in two different states.

We show only JGND and JY in Table 6.1. These are almost equal to each other in every
case. It is observed that NN radicals are much more strongly coupled to each other than the
IN radicals. This is due to the larger spin density on the carbon atoms of the O- N-C-N-O
fragments in NN diradicals. The spin density on the carbon atoms of N-C-N-O fragments in
IN-diradicals is much less.
The J value is greater for the –CF3 substituents than that for the −CH3 groups in the
closed form, and smaller in the open form. This is due to the bulkier group restricting the
angle of rotation (Φ) of the nitronyl nitroxide rings from the coupler plane. The average Φ
follows the orders 1 > 2 and 3 > 4. A smaller Φ gives a greater conjugation. The intra-ring
C−C distances are more or less same in the four closed species. Therefore, J exhibits the
reverse orders, 1 < 2 and 3 < 4. The calculated intra-ring C−C distance and dihedral angles
are given in Table 6.2.

94
Chapter 6 Photomagnetism

Table 6.1. Calculated exchange coupling constants (J) and total energies at UB3LYP/6-311+G(d,p)
level. The J values are calculated using the GND and Yamaguchi equations.

Species EB (a.u.)
B ET (a.u.) JGND JY
<S2> <S2> (cm−1) (cm−1)
Eq.(2.5) Eq.(2.7)
1a −1762.3338519 −1762.3338619 2.20 2.20
1.0670 2.0679
1b −1762.3569120 −1762.3569596 10.43 10.37
1.0750 2.0805

2a −2357.9438295 −2357.9438367 1.58 1.58


1.0700 2.0707
2b −2357.9639160 −2357.9639662 11.02 10.95
1.0784 2.0847

3a −1611.9511192 −1611.9511208 0.34 0.34


1.0727 2.0235
3b −1611.9723107 −1611.9723196 1.95 1.95
1.0242 2.0253

4a −2207.5623297 −2207.5623308 0.22 0.22


1.0243 2.0245
4b −2207.5831239 −2207.5831336 2.13 2.13
1.0250 2.0263

The opposite effect is found for the open form. The reason is that the substitution of a
bulkier group increases the intra-ring C−C distance by about 0.065 Å, and the bridging C−C
bond lengths also increase. This causes the phenyl rings that are no longer coplanar in CPD
to move further away from each other, thereby weakening the magnetic interaction.
The difference between the magnetic properties of a and b species are not due to the
angle Φ, as the average value of Φ always follows the order b > a. The stronger magnetism in
the b species is evidently an outcome of the shorter route for the transmission of magnetic
interaction and the planarity of the coupler.
The total energy difference between the a and b species in the triplet state are nearly
the same for 1-4. The ratio J b/J a is largest in case of 4, but species 2 is undoubtedly the best
photomagnetic molecule. The (Jb−Ja) for substituted dihydropyrenes is clearly much larger
than those for the diradicals based on perfluorocyclopentene.

95
Chapter 6 Photomagnetism

Table 6.2. Dihedral angles of all the species.

C-C Φ1 Φ2 Φ3 Φ4 Aveg.

1a 2.75976 20.611 26.604 42.221 49.334 34.69


1b 1.54620 22.861 25.359 53.361 53.361 38.74
2a 2.82490 15.687 22.062 39.431 47.673 31.21
2b 1.55642 20.849 23.193 52.864 53.218 37.53
3a 2.75685 14.143 22.947 46.774 53.872 34.43
3b 1.54673 28.965 33.980 51.331 56.277 42.64
4a 2.82046 13.367 22.806 41.856 50.216 32.06
4b 1.55761 26.887 31.878 51.676 56.243 41.67

Table 6.3. Calculated magnetic exchange interactions at UB3LYP/6-311+G(d,p) level.

Species EB (a.u.)
B ET (a.u.) JGND
<S2> <S2> (cm−1)
Eq.(2.4)
5a -3347.46082819 -3347.4608282 2.2 x10−3
1.0714 2.0714
5b -3347.43784677 -3347.43780447 -9.28
1.0785 2.0745

6a -4451.22202588 -4451.22202588 0.00


1.0832 2.0832
6b -4451.20213889 -4451.20213591 -0.65
1.0844 2.0841

7a -4566.63412348 -4566.63410879 -3.22


1.0867 2.0854
7b -4566.58452280 -4566.58452235 -0.09
1.0867 2.0867

8a -2281.81941755 -2281.81941770 0.033


1.0687 2.0687
8b -2281.73263186 -2281.73271138 17.45
1.0780 2.0904

9a -3292.34823490 -3292.34823520 0.07


1.0713 2.0713
9b -3292.31447756 -3292.31437285 -22.98
1.0722 2.0597

96
Chapter 6 Photomagnetism

F2 F2
F2 F2 F2 F2

_ _
O O O O
UV
+N S S N +N S S N
Vis
N _ N N _ N
O O + O O +

5a 5b

F2 F2
F2 F2 F2 F2

_ O UV _ O
O S S O S S
N Vis N
+N S S +N S S

N _ N
N _
O + N O +
O O

6a 6b

F2 F2 F2 F2

F2 F2
_
O S O O S O
+ +
N N UV N N

N S S S N Vis N S S S N
O _O + O
_
O
7a 7b

_ UV
O O _ O
Vis O
N _ N
+N O _
O +N O O
N N
N + N +

8a 8b
NC CN NC CN
NC CN NC CN

CF3 CF3
UV
_ O _ O
O F3C O F3C
Vis
_ N _ N
+N O O +N O O N
N
N + N +

9a 9a

Figure 6.4. A few recently synthesized (5-7) and a few predicted photomagnetic molecules (8-9).

A few recently synthesized photomagnetic molecules (5-7)2-4 and additional one (8-9)
are also investigated (Figure 6.4). These molecules are quite large and their molecular
geometries are optimized at ROHF/6-31G(d,p) level. The calculated J values at UB3LYP/6-
311+G(d,p) level and given in Table 6.3.

97
Chapter 6 Photomagnetism

The diradicals 5 shows a very interesting phenomenon. The open ring structure
shows a very weak ferromagnetic interaction whereas the closed ring structure (5b) shows
quite high antiferromagnetic interaction.
The spacer in species 6 is quite large. This is why the unpaired electrons in 6a do not
interact at al. In close ring species 6b, the interaction is weakly antiferromagnetic in nature.
The switching behavior of 7 is different than the rest of species studied in this Chapter.
The open ring is the “ON” state here and the closed one is “OFF” state, whereas the reverse is
true for the rest of the cases. This is because the conjugation is affected due to the Diels-alder
reaction. The species 5-7 were synthesized and the magnetic properties were studies by EPR
only. The photomagnetic properties of 8 and 9 are predicted here.

6.3. Conclusions
In conclusion, we predict that species 1, 2, 8 and 9 would be good photomagnetic
molecules with J varying by a few cm−1 upon irradiation. Besides, these species are all
ferromagnetically coupled. The a forms (CPD) have very small singlet-triplet energy
differences and would be faintly magnetic. The b isomers (DDP), however, would retain a
fairly considerable magnetic character at a low temperature, and possibly also in an inert
matrix.

6.4. References
1. (a)Tanifuji, N.; Matsuda, K.; Irie, M. Polyhedron 2005, 24, 2484. (b) Matsuda, K.; Irie, M. Polyhedron
2005, 24, 2477. (c) Matsuda, K. Bull. Chem. Soc. Jap. 2005, 78, 383. (d) Tanifuji, N.; Matsuda, K.;
T

Irie, M. Org. Lett. 2005, 7, 3777. (e) Matsuda, K.; Irie, M. J. Photochem. Photobio. C: Photochem.
Rev. 2004, 5, 69.
2. Tanifuji, N.; Irie, M.; Matsuda, K. J. Am. Chem. Soc. 2005, 127, 13344.
3. Matsuda, K.; Matsuo, M.; Irie, M. J. Org. Chem. 2001, 66, 8799.
4. Matsuda, K.; Irie, M. Chem. Lett . 2000, 16.
5. Teki, Y.; Toichi, T.; Nakajima, S. Chem. Eur. J. 2006, 12, 2329.
6. Huai, P.; Shimoi, Y.; Abe ; S. Phys. Rev. B 2005, 72, 094413.
7. Mitchell, R. H.; Ward, T. R.; Chen, Y.; Wang, Y.; Weerawarna, S. A.; Dibble, P. W.; Marsella, M. J.;
Almutairi, A.; Wang, Z.-Q. J. Am. Chem. Soc. 2003, 125, 2974.
8. Williams, R. V.; Edwards, W. D.; Mitchell, R. H.; Robinson, S. G. J. Am. Chem. Soc. 2005, 127, 16207.
9. Trindle, C.; Datta, S. N. Int. J. Quantum Chem. 1996, 57, 781. (b)Trindle, C.; Datta, S. N.; Mallik, B. J.
Am. Chem. Soc. 1997, 119, 12947.

98
Chapter 7
Dinuclear Copper Complex

This Chapter describes the investigation of the magnetic properties of a recently


synthesized dinuclear complex, [Cu2(μ−OAc)4(MeNHpy)2]. We have explicitly calculated
the overlap integral Sab between the two magnetic orbitals, and found a value of 0.8589.
Deviating from the common practice of replacing Sab by 1, the computed value of the
integral has been used in calculating the magnetic exchange coupling constant (J). The
calculated J is −290 cm−1, in excellent agreement with the observed value of −285 cm−1.
Also, the calculated J value is a weakly varying function of the Cu-Cu distance.
Furthermore, we have shown that the onset of intramolecular hydrogen bonding reduces the
spin density on the bridging atoms and consequently the magnitude of J. This explains why
the complex under investigation has a J value smaller than that of [Cu2(μ−OAc)4(H2O)2]
(−299 cm−1). While establishing this trend, we predict that the complex [Cu2(μ−OAc)4(py)2]
would have a higher J value, about −300 cm−1.
Chapter 7 Dinuclear Copper Complexes

7.1. Introduction
The study of intramolecular magnetic coupling between two metal magnetic centers
within a molecule is a fascinating subject.1 Intramolecular and intermolecular magnetic
interactions play the major role in controlling the magnetic properties of molecular crystals.
The intramolecular magnetic coupling in dinuclear transition metal complexes is controlled by
the number of bridging ligands, the angle M(metal)−L(bridging-ligand)−M(Metal), the M−M
distance, and the nature of bridging ligands and other ligands. A large body of theoretical and
experimental work has been performed to explain the magnetic properties.2 A number of
compounds with the same basic structure Cu2(μ-OAc)4 have been synthesized and the
cooperative magnetic interactions in these have been investigated, thereby establishing that
the through-ligand superexchange leads to the rather strong intramolecular interaction in these
complexes.3−6 Recently, Barquín et al.7 have synthesized a similar compound by introducing
2-methyl imino pyridine in the axial position as shown in Figure 7.1. This has the important
characteristic of having two intramolecular hydrogen bonds as shown by the dotted lines,
while the corresponding analogs with H2O and NH3 molecules as axial ligands do not have
intramolecular hydrogen bonds.

Complexes with the Cu2(μ-OAc)4 basic structure have quite high intramolecular magnetic
exchange coupling constants.8 The coupling constant J is –299 cm–1 for [Cu2(μ-
OAc)4(H2O)2] and −285 cm−1 for [Cu2(μ-OAc)4(NH3)2]. The strong antiferromagnetic
interaction in these compounds was initially thought to be explained by the δ-overlap of the
dx2−y2 orbitals of the two copper atoms, which contain the unpaired electron.9 It is now
theoretically accepted that the through-space interaction has much less contribution to the
overall J in comparison to the through-bridge exchange contribution.10

Ruiz et al. have extensively carried out theoretical calculations by varying the substituents
of the bridging carbon atoms.10 These substitutents have been found to play a very important
role in controlling the value of J. As the electronegativity of the atom or the group attached to
the bridging carbon atoms increases, the J value generally decreases. In the case of very
strongly electronegative groups like −CF3 and −CCl3, however, the opposite effect is
observed. Ruiz et al.10a have also noticed that the changes in axial ligands introduce only a
little variation of J. The qualitative model of Hay-Thibeut-Hoffmann (HTH)11 cannot explain
the observed variation of J in acetate-bridged compounds. This model is mainly based on the
99
Chapter 7 Dinuclear Copper Complexes

Figure 7.1. The dinuclear copper complex [Cu2(μ−OAc)4(MeNHpy)2].

(a) (b)

Figure 7.2. Structure of (a) the pyridine substituted complex, and (b) a conformer
with the MeNHpy ligands in the axial positions rotated through 45°.

energy differences between the singly occupied molecular orbitals (SOMOs). The
participation of the axial ligands in the SOMOs is forbidden by symmetry. It is also generally
known that the coupling constant depends on the number of bridging ligands n and can be
expressed as J = JF + nJAF where JF and JAF are respectively the ferromagnetic and
antiferromagnetic contributions. In addition to these controlling factors, the spin density of
the bridging atoms must be ultimately responsible for the J value. Any physical phenomenon

100
Chapter 7 Dinuclear Copper Complexes

that can affect the spin density on the bridging atoms can control the intramolecular coupling
constant.

Table 7.1. The computed overlap integral between the spatial parts of alpha and beta occupied
orbitals in the BS state of the dinuclear copper complex.

MO Orbital Energy (a.u.) Saba


α β

HOMO −0.20278 −0.20273 0.9997


HOMO-1 −0.20294 −0.20299 0.9997
HOMO-2 −0.23769 −0.23753 −0.2876
HOMO-3 −0.24290 −0.24310 −0.4977
HOMO-4 −0.25369 −0.25370 0.7405
HOMO-5 −0.26021 −0.26016 0.7478
HOMO-6 −0.26385 −0.26384 0.9913
HOMO-7 −0.26912 −0.26914 −0.8126
HOMO-8 −0.27105 −0.27098 0.7020
HOMO-9 −0.27914 −0.27915 −0.9844
HOMO-10 −0.28208 −0.28206 0.9554

a
Overlap integral between the magnetic orbital not between the metal atoms.

The main objective of this work is to carry out a theoretical investigation and evaluate the
magnetic exchange coupling constant by broken-symmetry approach while avoiding any
approximation to the overlap integral between the magnetically active orbitals. We have
calculated the intramolecular magnetic exchange coupling constant for the complex
synthesized by Barquín et al.7 The second aim is to reinvestigate the mechanism of the
magnetic exchange interaction of the same complex. This has been done in two ways. First,
we have shown that the contribution of the overlap between the orbitals of the two copper
atoms to the overlap of the two magnetic orbitals is truly small. Second, we have performed
computations on structures with varying Cu-Cu distance while the rest of the structure is kept
intact, and analyzed the spin distribution in each case. The third goal here is to demonstrate
that the J value undergoes reduction with the onset of intramolecular hydrogen bonding. This
has been investigated by comparing the spin density distribution in three systems, namely, the
original complex, a complex with pyridine substitution in the axial position (a possible new
compound) and a complex with the axial ligands in a different conformation. The latter
structures are shown in Figure 7.2.

Table 7.2. Single-point calculations in UB3LYP method using 6-311G(d,p) basis sets for atoms
other than the Cu atoms.
101
Chapter 7 Dinuclear Copper Complexes

Cu basis set Energy (a.u.) EBS−ET Ja


<S2> (cm−1) (cm−1)
BS T

6-311G(d,p) −4881.0236208 −4881.0225427 −237 −272


0.987599 2.002953
Lanl2dz −4881.0934244 −4881.0922747 −252 −290
0.983606 2.003598
Cep-121G −4881.1188212 −4881.1176725 −252 −290
0.987053 2.003776

a
Using Sab= 0.859 in Eq. (2.2).

7.2. Computational Methodology

The molecular geometry of the complex is obtained from the crystal structure reported
by Barquín et al. in Ref 7. The geometry was not theoretically optimized, as we want to keep
up the effects of the surrounding molecules on the geometry of the selected molecule in the
real crystal. To obtain the broken-symmetry states, all the single-point UB3LYP calculations
have been performed with the accurate guess values of molecular orbitals, which are in turn
retrieved from the proper ROHF calculations. In all the calculations, 6-311G(d,p) basis sets
are used for the lighter atoms. For the copper atoms, we have used both LANL2DZ and CEP-
121G basis sets. We have ourselves written a program to compute the overlap integral Sab

between the two magnetic orbitals, and to evaluate the contribution of the overlap between the
orbitals of the two copper atoms to Sab . This program uses the basis set information and

molecular orbital coefficients from the log files of Gaussian 98 software.

7.3. Choice of Magnetic orbitals

The interaction between two magnetic centres A and B is similar to the existence of
very weak chemical bond between them. In the present case, both the singlet and triplet states
are very close to each other in energy, and both can be thermally populated. When the
interaction vanishes, the dynamics of the two active electrons (one from A and another from
B) become uncorrelated. The Heitler-London approach in this case results in two
semilocalized orbitals a and b which can be identified as the magnetic orbitals.

102
Chapter 7 Dinuclear Copper Complexes

Orthogonalized magnetic orbitals and natural magnetic orbitals are oft-quoted types of
magnetic orbitals.12 The computational procedure adopted for the BS calculations here
necessitates the choice of the natural magnetic orbitals.

Table 7.3. Single-point calculations by UB3LYP method using 6-311G(d,p) basis sets for
atoms other than the Cu atoms. The Cu-Cu distance is varied.

Cu basis Cu-Cu Energy (a.u.) EBS−ET Ja


set Distance(Å) <S2> (cm−1) (cm−1)
BS T
2.62 −4881.061151 −4881.0601861 −212 −244
0.981953 2.003702
2.67 −4881.083968 −4881.0828751 −240 −276
0.983567 2.003693
Lanl2dz 2.72 −4881.0934244 −4881.0922747 −252 −290
0.983606 2.003598
2.77 −4881.09279690.9 −4881.0916344 −255 −294
82964 2.003604
2.82 −4881.0844034 −4881.083226 −258 −297
0.982456 2.00372
2.87 −4881.065495 −4881.0643415 −253 −291
0.983909 2.003576

2.62 −4881.0870304 −4881.0859311 −241 −278


0.987054 2.003776
2.67 −4881.1097856 −4881.1086601 −247 −284
0.987016 2.003789
Cep-121g 2.72 −4881.1188212 −4881.1176725 −252 −290
0.987053 2.003776
2.77 −4881.1198784 −4881.1187121 −256 −295
0.986935 2.003775
2.82 −4881.1114843 −4881.110294 −261 −301
0.98607 2.003824
2.87 −4881.0924631 −4881.0913333 −248 −285
0.987672 2.003771

a
Using Sab= 0.859 in Eq. (2.2).

The overlap integral between the α-HOMO and β-HOMO in the BS state is nearly
unity (0.9993) (Table 7.1). The calculated expectation value of S2 for the BS states are
approximately equals to 0.98 (Table 7.2), indicating that these HOMOs cannot be considered
as the magnetic orbitals. It is well known that the magnetic properties of transition metal
complexes evolve from the nature of the orbitals and their electronic population in the
transition metal ions. Therefore, it is reasonable to expect that the magnetic orbitals are to be
determined by the copper atoms.

103
Chapter 7 Dinuclear Copper Complexes

To obtain the natural magnetic orbital a for the present complex


[Cu(7)(μ−OAc)4Cu(13)(MeNHpy)2], single-point calculations have been performed by
treating Cu(13) as dummy to prevent any orbital interaction between Cu(13) and the rest of
the molecule. The resulting singly-occupied HOMO is chosen as orbital a. Similarly, the
magnetic orbital b has been obtained by treating atom Cu(7) as dummy. These calculations
have been carried out at the UB3LYP/6-311G(d,p) level. The computed overlap integral
between the selected magnetic orbitals is 0.8589(49).

-4881.06
Energy (a.u.)

-4881.12

2.600 2.925
Cu-Cu distance in Angstrom

Figure 7.3. The plot of the calculated total energy against the Cu-Cu distance. The total energy
was calculated by UB3LYP method with 6-311G(d,p) basis sets for atoms other than
the Cu atoms. The dotted lines are for the triplet states and bold lines are for the
broken-symmetry states. The upper set of curves is for the lanl2dz basis set used for
the copper atoms and the lower one is for the CEP-121G basis set.

7.4. Results and Discussion


The intramolecular magnetic coupling constants are calculated using Eq. (2.2). The
computed value of the overlap integral Sab remains more or less unchanged through three
digits for the different basis sets and the varying Cu-Cu distance investigated here. Therefore,

104
Chapter 7 Dinuclear Copper Complexes

Sab= 0.859 has been used throughout in these calculations to generate a three digit accuracy
for the

Table 7.4. The variation of average spin density square and the difference of average spin density
square between the Broken-Symmetry and Triple state for Cu(7) and Cu(13).

Cu-Cu 2.62 2.67 2.72 2.77 2.82 2.87


distance (Å)

P2HS(Cu)
P 0.4533 0.4568 0.4585 0.4637 0.4643 0.4703
P2BS(Cu)
P 0.4497 0.4526 0.4536 0.4566 0.4577 0.4641
-ΔP2(Cu) P
−0.0037 −0.0043 −0.0049 −0.0071 −0.0066 −0.0063

J (cm-1) −278 −284 −290 −295 −301 −285

calculated J. The calculated J values are given in Table 7.2. The metal atoms are known to
control the magnetic properties of transition metal complexes, and it is necessary to treat these
atoms with effective core potentials. In fact, when we use the 6-311G(d,p) basis set for the
valence orbitals of copper atoms, we obtain a J value of –272 cm−1 from Eq. (2.2), whereas
both CEP-121G and LANL2DZ bases give a coupling constant of −290 cm−1. The latter J
values excellently match the observed value –285 cm−1 that was reported by Barquín et al.7
The better agreement for LANL2DZ basis set is largely fortuitous, as the LANL2DZ basis set
is not necessarily better than the 6-311G(d,p) basis. But the improved value obtained from
the use of CEP-121G basis is notable.

The overlap of the d-orbitals of the two copper atoms (mainly dx2−y2) in the magnetic
orbitals is indicative of δ-bonding between the two Cu atoms, that is, the strength of the
through-space interaction. The computed value of this overlap integral is around 0.000011 in
absolute magnitude. This is not surprising as the two copper atoms are considerably away
from each other, but it confirms that the direct exchange would be negligibly small. In each
complex with one copper atom as dummy, the gross d-orbital population is 8.75 and the total
electronic population is 27.32 for the remaining copper atom. To test the absence of direct
exchange further, the distance between the two copper atoms is varied from 2.62 Å to 2.87 Å
with an interval of 0.05 Å. The rest of the structure is kept intact at the crystallographic
geometry. The observed Cu-Cu distance in the crystallized complex is 2.72 Å. Table 7.3
shows a very systematic variation of the single-point total energy with the Cu-Cu distance.

105
Chapter 7 Dinuclear Copper Complexes

Figure 7.3 illustrates this trend. The CEP-121G basis set consistently yields a lower energy
for each state. Both the bases yield a more stable BS configuration. In the case of the

Table 7.5. Single-point calculations by UB3LYP method using 6-311G(d,p) basis sets for the
atoms other than Cu atoms by replacing the axial ligands with pyridine and by rotating the axial ligands
through 45º.

Cu basis Condition Energy (a.u.) EBS−ET Ja


set <S2> (cm−1) (cm−1)
BS T

Original −4881.0934244 −4881.0922747 −252 −290


0.983606 2.003598
Lanl2dz Pyridine −4691.7378766 −4691.7366646 −266 −306
subst. 0.982556 2.003666
45º rotation −4881.0903817 −4881.08917932.0 −264 −303
0.985857 03708

Original −4881.1188212 −4881.1176725 −252 −290


0.987053 2.003776
Cep-121G Pyridine −4691.7654575 −4691.764268 −261 −300
subst. 0.986742 2.003831
45º rotation −4881.1147885 −4881.1136113 −258 −297
0.986682 2.00381

a
Using Sab= 0.859 in Eq. (2.2).

LANL2DZ basis set, the minimum of the energy curve is around 2.74 Å. For CEP-121G basis
set it is at about 2.75 Å. These values are in good agreement with the observed Cu-Cu
distance of 2.72 Å in crystal. We notice from Table 7.3 that the calculated J value is only
weakly dependent on the Cu-Cu distance. The absolute magnitude of the calculated coupling
constant increases as the distance increases up to about 2.82 Å and then it decreases (Figure
7.4). This is only possible if there is no δ-bonding. The through-space interaction contributes
very little to the J value, and the through-bridge exchange is the dominant contribution. The
slight variation of the J value arises from the change in the overlap and bonding with the
atoms of the bridging ligands.
The reason for the absence of direct exchange can be understood from Figure 7.5. The
d x 2 − y2 orbitals of the copper atoms as shown in Figure 7.5(b) form the singly occupied

106
Chapter 7 Dinuclear Copper Complexes

HOMO’s whereas the d z 2 orbitals in Figure 7.5(e) are fully occupied. As the d x 2 − y2 point

towards the bridging ligands while the d z 2 orbitals are along the Cu-Cu axis, the through-

bridge interaction determines the magnetic properties. The overlap between the d z 2 orbitals

is negligibly small.

Table 7.6. The calculated spin densities on different atoms from the UB3LYP/6-311G(d,p) + Cep-
121G results for the broken-symmetry sate.

Cu(7) O(8) C(9) O(31) Cu(13)


Original −0.6766 −0.0795 0.0004 0.0797 0.6705
Pyridine −0.6691 −0.0778 −0.0001 0.0787 0.6685
45° rotated −0.6736 −0.0765 0.0003 0.0766 0.6687

Cu(7) O(30) C(20) O(28) Cu(13)


Original −0.6766 −0.0785 0.0002 0.0805 0.6705
Pyridine −0.6691 −0.0785 0.0001 0.0779 0.6685
45° rotated −0.6736 −0.0758 0.0002 0.0775 0.6687

Cu(7) O(34)a C(26) O(25) Cu(13)


Original −0.6766 −0.0650 0.0660 0.0804 0.6705
Pyridine −0.6691 −0.0721 -0.0011 0.0802 0.6685
45° rotated −0.6736 −0.0707 -0.0010 0.0829 0.6687

Cu(7) O(24) C(23) O(22)a Cu(13)


Original −0.6765 −0.0791 0.0018 0.0660 0.6705
Pyridine −0.6691 −0.0803 0.0011 0.0719 0.6685
45° rotated −0.6736 −0.0823 0.0014 0.0718 0.6687

a
Hydrogen bonded oxygen atoms in original compound.

The average squared spin populations of the Cu atoms are correlated with the
computed J values in Table 7.4. The spin densities are obtained from the calculation at
UB3LYP/6-311G(d,p)/CEP-121G level. The variation of −ΔP2(Cu)
P [defined as
ΔP2(Cu)=P2(Cu)HS−P2(Cu)BS] with the Cu-Cu distance (Table 7.4) is in general agreement
P P

with the nature of the plot of the calculated J versus the Cu-Cu distance (Figure 7.4). But for
the O atoms, it is observed that only the triplet spin population matches with the nature of
variation of J.
The intramolecular hydrogen bonding has a minor influence on the magnetic exchange
coupling constant. Table 7.5 shows that the absolute magnitude of J increases by about 10
cm–1 for the pyridine substituted species and for the 45º-rotated conformation. To find a

107
Chapter 7 Dinuclear Copper Complexes

reason for this behavior, we have investigated the spin density on the bridging oxygen atoms.
The spin density distribution is shown in Table 7.6. The spin density changes progressively
along each Cu-O-C-O-Cu chain, manifesting an antiferromagnetic trend. The spin densities
on the hydrogen-bonded O(22) and O(34) atoms in the original complex are lower than those
on the non-hydrogen-bonded oxygen atoms, and also much reduced compared to the spin
densities in the pyridine substituted complex and the 45º–rotated conformer. The spin
densities on all other oxygen atoms remain almost unchanged in the latter two species. Thus
the intramolecular hydrogen bonding reduces the spin density on the hydrogen bonded
oxygen atoms. This reduction leads to a diminished extent of the through-bridge magnetic
interaction, thereby lowering the absolute magnitude of the J value.

-220
J (cm -1)

-320
2.6 2.7 2.8 2.9
Cu-Cu distance in Angstrom

Figure 7.4. Variation of the calculated J values with the Cu-Cu distance for the lanl2dz (dotted
line) and CEP-121G (bold line) copper basis sets.

The complex [Cu2(μ−OAc)4(py)2] is seen to have a J value of about −300 cm−1 (Table 7.5)
that is as much as that for the water-substituted complex.

In molecular magnetism, the influence of hydrogen bonding on the spin-spin interaction


and spin migration is a very common phenomenon.13 Recently, Desplanches et al.14 have
reported a computational study on dinuclear Cu(II) complexes with two monomeric units
linked by O−H···O to form a dimer. These authors noted that the hydrogen-bonded H atom
does not have a major contribution to the SOMOs but it takes part in spin density

108
Chapter 7 Dinuclear Copper Complexes

delocalization between the two Cu atoms. This leads to an intramolecular magnetic

(a).
(b)

(c)

(e)
(d)

Figure 7.5. View of molecular orbitals: (a) LUMO, (b) HOMO (c) HOMO-1 (d) HOMO-2, 109
and (e) HOMO-5. The electronic population of HOMO and HOMO-1 is 1.0.
The MOs are obtained from the ROHF/6-311G(d,p) /Lanl2dz calculation for
the triplet.
Chapter 7 Dinuclear Copper Complexes

coupling. In the present case, we notice the opposite phenomenon, that is, the spin density
delocalization in the superexchange pathway is reduced by the formation of a hydrogen bond.
The difference is that the monomeric units are directly linked here and hydrogen bonding
reduces the spin distribution in this linkage.

7.5. Conclusions

The magnetic exchange coupling constant of the recently synthesized dinuclear copper
complex [Cu2(μ-OAc)4(MeNHpy)2] has been calculated by using broken-symmetry density
functional methodology, and in doing so we have explicitly computed the overlap integral
between the two magnetic orbitals. The α-HOMO and β-HOMO in BS state are not magnetic
orbitals. The magnetic orbitals are obtained by using the concept of natural magnetic orbitals.
The calculated magnetic exchange coupling constant −290 cm−1 is in good agreement with the
observed value of −285 cm−1.

The direct exchange between the two copper atoms is negligibly small, and the
superexchange interaction is predominant. This conclusion is made after determining the
contribution of the overlap between the orbitals of the two copper atoms to Sab, and also by
studying the spin density distribution while the Cu-Cu distance is varied. There is a lack of δ-
bonding between the d x 2 − y2 orbitals that carry the unpaired electrons. Intramolecular

hydrogen bonding reduces the spin density of the oxygen atoms, and leads to a lower absolute
magnitude of J as compared to the complex that contains H2O instead of MeNHpy. The
magnitude of the J value for [Cu2(μ−OAc)4(py)2] will be as high as that for the complex with
water as axial ligands.

7.6. Reference
1 (a) Kahn, O. Molecular Magnetism; VCH: New York, 1993. (b) Goodenough J. B. Magnetism and the
Chemical Bond; Interscience: New York, 1963. (c) Coronado, E.; Delhaè, P.; Gatteschi, D.; Miller, J. S.
Molecular Magnetism: From Molecular Assemblies to the Devices, Eds.; Nato ASI Series E, Applied
Sciences, Kluwer Academic Publisher: Dordrecht, Netherland, 1996; Vol. 321 (d) Benelli, C.; Gatteschi,
D. Chem. Rev. 2002, 102, 2369. (e) Millar, J. S.; Drillon, M. Magnetism: Molecules to Materials
Nanosized Magnetic Materials, Wiley-VCH, Weinheim, 2002.
2 (a) Master, P. De; Fletcher, S. R.; Skapski, A.C. J. Chem. Soc. Dalton Trans. 1973, 2575. (b) Catterick. J.;
Thornton, P. Adv. Inorg. Chem. Radiochem. 1977, 20, 291. (c) Doednes, R. J. Prog. Inorg. Chem. 1976,

110
Chapter 7 Dinuclear Copper Complexes

21, 209. (d) Rao, V. M.; Sathyanarayana, D. N.; Manohar, H. J. Chem. Soc. Dalton Trans. 1983, 2167. (e)
Nakagwa, M.; Inomata, Y.; Howell, F.S. Inorg. Chim. Acta 1999, 295, 121. (d) Ruiz, E.; Rodríguez-
Fortea, A.; Cano, J.; Alvarez, S.; Alemany, P. J. Comput. Chem. 2003, 24, 982. (e) Ruiz, E.; Rodríguez-
Fortea, A.; Alvarez, S. Inorg. Chem. 2003, 42, 4881.
3 Guha, B. Proc. R. Soc. 1952, 206, 353.
4 Bleaney, B.; Bowers, K. D. Proc. R. Soc. London, A, 1952, 214, 451.
5 Niekerk, J. N. van; Schoening, F. R. L. Acta. Crystallogr. 1953, 6, 227.
6 (a) Jotham, R. W.; Kettle, S. F. A.; Marks, J. A.; J. Chem. Soc. Dalton Trans. 1972, 428. (b) Melníc, M.
Coord. Chem. Rev. 1981, 36, 1. (c) Muto, Y.; Nakashima, M.; Tokii, T.; Kato, M.; Suzuki, I. Bull. Chem.
Soc. Jpn. 1987, 60, 2849. (d) Steward, O. M.; McAfee, R. C.; Chang, S.-C.; Piskor, S. R.; Schreiber, W. J.;
Jury, C. F.; Taylor, C. E.; Pletcher, J. F.; Chen. C.-S. Inorg. Chem. 1986, 25, 771.
7 Barquín, M; Garmendia, M. J. G.; Pacheo, S.; Pinilla, E.; Quintela, S.; Seco, J. M., Torres, M. R. Inorg.
Chim. Acta 2004, 357, 3230
8 (a) Sesco, J. M.; González Garmendia, M. J.; Pinilla, E.; Torres, M. R. Polyhedron 2002, 21, 457. (b)
Jotham, R. W.; Kettle, Sidney F. A. Chem. Comm. 1969, 6, 258. (c) Sesto, R. E. D.; Deakin, L; Miller, J.
S. Synthetic Metals 2001, 122, 543. (d) Figgis, B. N.; Martin, R. L. J. Chem. Soc. 1956, 3837. (e)
Yablokov, Yu. V.; Mosina, L. V.; Simonov, Yu. A.; Milkova, L. N.; Ablov, A. V.; Ivanov, V. I. Zh.
Strukt. Chim. 1978, 19, 42.
9 (a) Datta, R. L.; Syamal, A. Elements of Magnetochemistry; Affiliated Ease-West Press Pvt. Ltd.: New
Delhi, 1993. (b) MaxDougall, J. J.; Nathan, L. C.; Nelson, J. H. Inorg. Chim. Acta 1976, 17, 243.
10 (a) Rodríguez-Fortea, A.; Alemany, P.; Alvarez, S.; Ruiz, E. Chem. Eur. J. 2001, 7, 627 (b) Rodríguez-
Fortea, A.; Alemany, P.; Alvarez, S.; Ruiz, E. Inorg. Chem. 2002, 41, 3769. (c) Ruiz, E.; Llunell, M.;
Alemany, P. Sold. Stat. Phys. 2003, 176, 400.
11 Hay, P. J.; Thibeault, C. J.; Hoffmann, R. J. Am. Chem. Soc. 1975, 97, 4884.
12 R. D. Willett, D. Gatteschi and O. Kahn, Magneto-Structural Correlations in Exchange Coupled
Systems, NATO ASI Series, Reidel, Dordrecht, 1985.
13 (a) Romero, F. M.; Ziessel, R.; Bonnet, M.; Pontillon, Y.; Ressouche, E.; Schweizer, J.; Delley, B.; Grand,
A.; Paulsen, C. J. Am. Chem. Soc. 2000, 122, 1298. (b) Hicks, R. G; Lemaire, M. T.; Öhrström, L.;
Richardson, J. F.; Thompson, L. K.; Xu, Z. J. Am. Chem. Soc. 2001, 123, 7154.
14 Desplanches, C.; Ruiz, E.; Rodríguez-Fortea, A.; Alvarez, S., J. Am. Chem. Soc. 2002, 124, 5197.

111
Conclusions

Conclusions

The magnetic exchange coupling constant J, that is related to the difference between
the energy of singlet (ES) and triplet (ET) states, is theoretically examined. A theoretical
formalism on the N-electron spin Hamiltonian is given. From this treatment, a theoretical
expression for J is easily derived, which can be reduced to the popular spin projection formula
due to Giensberg, Noodleman and Davidsion. A few initial calculations are done by
employing post-Hartree-Fock methods. The rest of the calculations are based on the density
functional, broken symmetry methodology. The broken-symmetry calculations are easy to
perfom even on larger molecules using moderate computational facility.
The coupling constant J is very small, of the order of 10-5 a.u., whereas the energy
calculated for each spin state can differ from the accurate energy by a large amount. The
calculation of J relies on the fact that while both ES and ET (or ES and ET) can be in errors,
their difference is correctly reproduced by dependable methodology and a large basis set.
The effect of molecular vibration on J is not studied here. This is because the
vibrational frequencies are more or less the same in the two spin states in most cases. The
difference in vibrational energy would be large when the singlet and triplet geometries highly
differ from each other. The BS calculations are performed using the triplet geometries.
Calculations on a total 79 molecules are presented in this thesis. Each species is
investigated in two different spin state by a number of methodologies and using different
basis sets. A plethora of optimized bond angles, bond lengths and MO coefficients is
available from the respective journal sites where the parts of the work reported herein have
been published.
Fused ring diradicals are investigated in HF and DFT methodologies. The unrestricted
calculations produced very high spin contamination. The rule of spin alternation can predict
the ground spin state. The tautomeric conversion of H atom between singlet and triplet
optimized geometries is observed in a few diradicals.
Series of bis-nitrotronyl nitroxide diradicals with different conjugated couplers have
been investigated. The computed magnetic exchange coupling constants are in very good
agreement with the reported values. The α-HOMO and β-HOMO in the BS state are

112
Conclusions

generally found to be magnetic orbitals. In conjugated systems, the magnetic interaction


is mainly transmitted through the π-electron conjugation. The strength of antiferromagnetic
interaction decreases with the increase in the length of conjugated couplers. Conjugated
linear couplers are more efficient antiferromagnetic couplers than the aromatic ones of similar
length. The diradicals with m-couplers are undoubtedly ferromagnetic. The shape of the
SOMOs as well as the rule of spin alternation in the UHF emerge as two robust guidelines for
the prediction of the qualitative nature of the intramolecular magnetic interaction in bis-
nitronyl nitroxide diradicals. The calculated J also depends on the basis sets.
The coupling constant J is found to decrease for the linear acene couplers from one to
three benzenoid rings, but it increases from three to five benzenoid rings. The α-HOMO and
β-HOMO are not the only magnetically active orbitals for the molecules with 3-5 rings. This
happens due to the increase of the diradical character of the acene couplers. The diradical
character is lost in the bent couplers. The NICS value at the central rings of the linear acene
is high, while the terminal rings lose some of the benzenoid character. The J value increases
with the bond order, and decreases with the increase in the angle of twist of the NN mono-
radicals from the coupler plane. The qualitatively proposed equation (5.1) can give a fair
estimate of J for these molecules. Reliable aN values are obtained for the diradicals in
solution.
We predict a few species that would be good photomagnetic switch molecules with J
varying by a few cm−1 upon irradiation. The predicted species are all ferromagnetically
coupled.
The magnetic exchange coupling constant of the recently synthesized dinuclear copper
complex [Cu2(μ-OAc)4(MeNHpy)2] has been calculated by using broken-symmetry density
functional methodology, and in doing so we have explicitly computed the overlap integral
between the two magnetic orbitals. The complex has several pairs of magnetic orbitals
although HOMO and HOMO-1 pairs are not magnetic orbitals. Overlap integrals are
computed by using the concept of natural magnetic orbitals. The calculated magnetic coupling
constant −290 cm−1 is in good agreement with the observed value of −285 cm−1. The direct
exchange between the two copper atoms is negligibly small, and the superexchange
interaction is predominant. This conclusion is made after determining the contribution of the
overlap between the orbitals of the two copper atoms to Sab, and also by studying the spin

113
Conclusions

density distribution while the Cu-Cu distance is varied. There is a lack of δ-bonding between
the d x 2 − y2 orbitals that carry the unpaired electrons. The Bencini-Ruiz formula is applicable

to such systems due to the presence of highly degenerate HOMOs. Intramolecular hydrogen
bonding reduces the spin density of the oxygen atoms, and leads to a lower absolute
magnitude of J as compared to the complex that contains H2O instead of MeNHpy. The
magnitude of the J value for [Cu2(μ−OAc)4(py)2] is predicted to be as high as that for the
complex with water as axial ligands, nearly −300 cm−1.

114
Summary

Summary of the Thesis

The main aim of the thesis is to study the magnetic properties of diradicals, mostly of
organic origin. Molecule-based magnetic materials have several advantages over traditional
magnets. These are transparent, insulating, photoactive, thermally controllable and bio-active.
Molecular magnetism arises from the spin of the unpaired electron. Chapter 1 gives a general
introduction to magnetic molecules. The origin of magnetic phenomena in molecules and the
empirical ways of determination of the nature of magnetic exchange interaction are discussed.
The roles played by molecular magnetism in modern science and technology, such as
photomagnetic effect, single molecule magnets (SMM), spintronics etc., are also reviewed. The
objectives are clearly stated and a chapter-wise arrangement of the thesis is mentioned.
The theoretical background for the determination of magnetic exchange coupling constant
by broken symmetry (BS) calculations1 is investigated in Chapter 2. In this chapter we formulate
an N-electron spin Hamiltonian for diradicals having non-degenerate highest occupied molecular
orbitals. At first, energy expressions are obtained for singlet, broken-symmetry and triplet single-
determinant wave functions of unrestricted Hartree-Fock treatment. Total energy values for the
two-determinant singlet and triplet configurations that can be obtained from a self-consistent-
field treatment are determined next by using the orbital perturbation theory. This leads to an
energy ordering, which is expected to be valid also in an unrestricted Hartree-Fock Kohn-Sham
treatment. The base line of the spin Hamiltonian is determined from this ordering, and the spin
Hamiltonian is formulated. The spin Hamiltonian reduces to the Heisenberg effective spin
Hamiltonian operator in the two-center two-electron case. Using the spin Hamiltonian, we obtain
expressions for the average value of energy for the broken-symmetry and triplet determinants. It
is shown that the Yamaguchi expression2 for the magnetic exchange coupling constant J is
approximately valid. A more correct expression for J is based on spin projection. The latter
reduces to the Ginsberg-Noodleman-Davidson formula3 when the spin projection difference
equals ½. It normally happens in a density functional calculation using a large basis set. In the
latter case, a good estimate of the energy of the two-determinant singlet is also obtained. These

i
Summary

observations are briefly discussed by using examples of two nitronyl nitroxide diradicals, one
with a large, and the other with a very small, value of J.
In Chapter 3, we discuss our investigation of the ground state spins of seven diradicals
belonging to the fused ring system by traditional ab-initio methodologies. The systems under
study are (1) 4-oxy-2-naphthalenyl methyl, (2) 1,8-naphthalenediylbis(methyl), (3) 8-imino-1-
naphthalenyl methyl, (4) 1,8- naphthalenediylbis(amidogen), (5) 8-methyl-1-naphthyl carbene,
(6) 8-methyl-1-naphthalenyl imidogen and (7) 8-methyl-1-naphthyl diazomethane (Figure 3.1).
Out of the seven molecules, only 1 was theoretically investigated earlier. To our knowledge, for
2-7, this work represents the first ab-initio investigation. A variety of basis sets has been
employed in these calculations. For each spin state, the molecular geometry has been fully
optimized at the unrestricted Hartree-Fock (UHF) level using the STO-3G, 4-31G, 6-311G(d) and
6-311G(d,p) basis sets. The UHF optimized geometries have been used for Møller-Plesset (MP)
and coupled cluster (CC) calculations as well as the density functional (UB3LYP) treatment.
Results in the unrestricted formalism have been given only at UHF and UB3LYP levels for the 6-
311G(d) basis. The UHF calculations yield an unrealistically large Singlet−Triplet (S−T)
splitting. Splittings calculated with different bases disagree seriously. The S−T gap is smaller in
the split-valence bases. The basis set truncation error can be considerably overcome by
considering electron correlation. Møller-Plesset perturbation theory and UCCSD(T) does not
produce realistic S−T energy gap unless the basis set used is very large which imposes the limit
of computing ability.
For these diradicals, any meaningful result would require larger bases with polarization
functions. Apart form this difficulty, the optimized molecular geometries turned out to be highly
spin-contaminated. The spin-contamination can be significantly reduced by the density
functional UB3LYP treatment. Nevertheless, for most of the diradicals, the UB3LYP method did
not yield a systematic trend. To avoid spin contamination completely, we have repeated
computations in the restricted (open-shell) Hartree-Fock framework. Geometry optimizations
were carried out using STO-3G, 6-311G(d), and 6-311G(d,p) bases at R(O)HF level and 6-
311G(d,p) basis at R(O)B3LYP level for each spin state. The R(O)B3LYP/6-311G(d,p)
optimized geometry yields the best total energy for each spin state and hence the most reliable
ii
Summary

S−T energy difference. Molecules 1-6 are found as ground state triplets. The calculated results
are in agreement with the available experimental findings. Molecules 3 and 7 have widely
different geometries in the singlet and triplet states. The calculations using 6-311G(d) and 6-
311G(d,p) basis sets show that in molecule 3, the substituents of naphthalene are −NH2 and −CH
in singlet but −NH and −CH2 in triplet. The two optimized geometries are tautomeric forms.
Molecule 7 is expected to be either a ground state triplet with a very little S−T gap or a ground
state singlet. This prediction is borne out by the computed results. The R(O)B3LYP/6-
311G(d,p) calculation yields a S−T splitting of −21.9 kcal mol-1. The singlet state becomes
stabilized by forming an additional condensed ring. The UHF spin density plots obtained from
the 4-31G optimized geometries manifest the phenomenon of spin alternation in the ground state.
A series of Nitronyl Nitroxide (NN) diradicals with linear conjugated couplers and
another series with aromatic couplers have been investigated by broken-symmetry (BS) DFT
approach. These are shown in Figure 4.1. The results are discussed in Chapter 4. First we show
that an ethylenic coupler provides a very strong intramolecular magnetic interaction. A recently
synthesized nitronyl nitroxide derivative, D-NIT2, is investigated by ab initio quantum chemical
methods. The broken symmetry approach yields a coupling constant −541 K that is in good
agreement with the observed value in solid state. The overlap integral between the magnetically
active orbitals in the BS state has been explicitly computed and used for the evaluation of the
magnetic exchange coupling constant (J). The calculated J values are in good agreement with the
observed values in literature. The magnitude of J depends on the length of the coupler as well as
the conformation of the radical units. The aromaticity of the spacer decreases the strength of the
exchange coupling constant. The SOMO-SOMO energy splitting analysis where SOMO stands
for the singly-occupied molecular orbital, and the calculation of electron paramagnetic resonance
(EPR) parameters have also been carried out. The computed hyperfine coupling constants
support the intramolecular magnetic interactions. The nature of magnetic exchange coupling
constant can also be predicted from the shape of the SOMOs as well as the spin alternation rule in
the unrestricted Hartree-Fock (UHF) treatment. It is found that π-conjugation along with spin-
polarization play the major role in controlling the magnitude and sign of the coupling constant.
Moreover, the magnetic properties of the diradicals (2,2'-(1,2-ethynediyldi-4,1-
iii
Summary

phenylene)bis[4,4,5,5-tetramethyl-4,5-dihydro-1 H-imidozolyl-oxyl] (IN-2p-IN), 2,2'-(1,2-


ethynediyldi-4,1 3,1-phenylene)bis[4,4,5,5-tetramethyl-4,5-dihydro-1 H-imidozolyl-oxyl] (IN-
pm-IN) and 2,2'-(1,2-ethynediyldi-4,1 3,1-phenylene)bis[4,4,5,5-tetramethyl-4,5-dihydro-1 H-
imidazole-1-oxyl-3-oxide] (NN-pm-NN) are also investigated. The rule of spin alternation in the
UHF clearly shows that the radical sites are antiferromagnetically coupled in IN-2p-IN and
ferromagnetically coupled in NN-2p-NN and NN-pm-NN, in agreement with a previously
carried out experiment.
The molecular geometries are optimized at Hartree-Fock levels. This is followed by
single-point calculations using the density functional (UB3LYP) treatment and the
multiconfigurational (CASSCF) methodology. Magnetic exchange coupling constants are
determined from the broken symmetry approach. The calculated J values, −3.60 cm−1 for IN-2p-
IN, 0.16 cm−1 for NN-2p-NN and 0.67 cm−1 for NN-pm-NN, are in excellent agreement with the
observed values.
Chapter 5 describes our prediction of the intramolecular magnetic exchange coupling
constant (J) for eleven nitronyl nitroxide diradicals (NN) with different linear and angular
polyacene couplers. These are shown in Figure 5.1. For the linear acene couplers, J initially
decreases with increase in the number of fused rings. But from anthracene coupler onwards, the
J value increases with the number of benzenoid rings due to an increasing diradical character of
the coupler moiety. The J value for the diradical with a fused bent coupler is always found to be
smaller than that for a diradical with a linear coupler of the same size. Nuclear independent
chemical shift (NICS) is calculated and it is observed that the average of the NICS values per
benzenoid ring in the diradical is less than that in the normal polyacene molecule. An empirical
formula for the magnetic exchange coupling constant of a NN diradical with an aromatic spacer
is obtained by combining the Wiberg bond order (BO), the angle of twist (φ) of the monoradical
(NN) plane from the plane of the coupler, and the NICS values. A comparison of the formula
with computed values reveals that from tetracene onwards, the diradical nature of the linear acene
couplers becomes prominent thereby leading to an increase in the ferromagnetic coupling
constant. Isotropic hyperfine coupling constants are calculated by using the polarized continuum
model for the diradicals in different solvents, and also for the species in vacuum.

iv
Summary

Chapter 5 also describes the prediction of the magnetic exchange coupling constant (J) of
m-phenylene based nitronyl nitroxide (NN) diradicals with nine different substituents in three
unique (common ortho, ortho-meta and common meta) positions on the coupler unit by from the
broken-symmetry density functional (BS-DFT) calculations. The molecules are shown in Figure
5.3. Substitutions at common ortho position always have greater angles of twist between the spin
source and the coupler units. When the twist angle is very large, a change of the intramolecular
interaction from ferromagnetic to antiferromagnetic is observed. In these cases, the coupler-NN
bond order (BO) becomes small due to a partial population of the σ* orbital. Substitution at the
common meta position of m-phenylene in the diradical has little steric and hydrogen bonding
effects. The effect of electron withdrawing power of the subtituent does not reveal any clear cut
expression except for the single atom substitution. In the latter case, an ortho substitution leads
to a decrease of J and a meta substitution increases J with a decreasing −I effect. The nucleus-
independent chemical shift (NICS) value is found to decrease from the corresponding mono-
substituted phenyl derivatives. The exchange coupling constant can be estimated from an
empirical equation when there is hardly any stereo-electronic or hydrogen bonding effects that
may change the twist angles.
The interesting magnetic phenomenon of photomagnetism is explored in Chapter 6. In
this Chapter, we predict the photo-switching magnetic properties of four substituted
dihydropyrenes from broken-symmetry calculations. (Figure 6.2 and Figure 6.4) The magnetic
exchange coupling constants differ up to 17 cm−1. The intramolecular exchange interactions are
ferromagnetic in nature. The calculated coupling constants are much larger than those reported
earlier for photomagnetic organic molecules.
In Chapter 7 of this thesis, we present our results of investigation of the magnetic
properties of a recently synthesized dinuclear complex, [Cu2(μ−OAc)4(MeNHpy)2], by broken-
symmetry (BS) density functional (DFT) methodology. While for the molecules investigated in
the previous chapters the Noodleman,1 Yamaguchi,2 and Ginsberg-Noodleman-Davidson formula
(GND)3 have been applicable, the Bencini-Ruiz formula gives a better description of J in
transition metal complex diradicals.4 The complex has several pairs of magnetic orbitals.
Therefore, we have explicitly calculated the overlap integral Sab between the two natural

v
Summary

magnetic orbitals, and found a value of 0.8589. Deviating from the common practice of
approximating Sab by 1 for the strongly delocalized systems, the computed value has been used in
calculating the magnetic exchange coupling constant (J) from the two electron-two orbital BS
model. The calculated J is −290 cm−1, in excellent agreement with the observed value of −285
cm−1. The contribution of the overlap between the orbitals of the two copper atoms to Sab is
negligibly small. Also, the calculated J value is a weakly varying function of the Cu-Cu distance.
The last two observations confirm that the through-ligand superexchange phenomenon is
responsible for the high magnetic exchange interaction in the Cu2(μ−OAc)4 complex(es).
Furthermore, we have shown that the onset of intramolecular hydrogen bonding reduces the spin
density on the bridging atoms and consequently the magnitude of J. This explains why the
complex under investigation has a J value smaller than that of [Cu2(μ−OAc)4(H2O)2] (−299
cm−1). While establishing this trend, we have predicted that the complex [Cu2(μ−OAc)4(py)2]
would have a higher J value, about −300 cm−1.

vi
List of Publications

Published

1. Ali, Md. E.; Datta, S. N. “Polyacene Spacers in Intramolecular Magnetic Coupling” J. Phys. Chem. A 2006,
110, 13232.

2. Ali, Md. E.; Datta, S. N. “Density Functional Theory of Prediction of Enhanced Photomagnetic Properties
of Nitronyl Nitroxide and Imino Nitroxide Diradicals with Substituted Dihydropyrene Couplers,” J. Phys.
Chem. A, 2006, 110, 10525.

3. Ali, Md. E.; Datta, S. N. “Theoretical Investigation of Magnetic Properties of a dinuclear copper complex
[Cu2(μ-OAc)4(MeNHpy)2],” J. Mol. Struc. :THEOCHEM, 2006, 775, 17.

4. Ali, Md. E.; Datta, S. N. “Broken-Symmetry DFT Investigation on bis-Nitronyl Nitroxide Diradicals:
Influence of Length and Aromaticity of Couplers” J. Phys. Chem. A 2006, 110, 2776.

5. Ali, Md. E.; Vyas, S.; Datta, S. N. “Ab Initio Quantum Chemical Investigation of Spin States of Some
Mono and Diradical Derivatives of Imino Nitroxide and Nitronyl Nitroxide” J. Phys. Chem. A 2005, 109,
6272.

6. Vyas, S.; Ali, Md. E.; Hossain, E.; Patwardhan, S.; Datta, S. N. “Theoretical Investigation of Intramolecular
Magnetic Interaction Through An Ethylenic Coupler” J. Phys. Chem. A 2005, 109, 4213.

7. Datta, S. N.; Jha, P. P.; Ali, Md. E. “Ab-initio Quantum Chemical Investigation of the Spin States of Some
Fused Ring Systems” J. Phys. Chem. A 2004, 108, 4087.

Submitted and ‘in preparation’


1. Ali, Md. E.; Misra, A.; Datta, S. N. “N-electron Spin Interpretation of Magnetic Exchange Interaction in
Broken-Symmetry approach”.

2. Ali, Md. E.; Hossain, E.; Datta, S. N. “Theoretical Investigation of Substituted m-Phenylene Spacers as
Ferromagnetic Couplers in Nitronyl Nitroxide Diradicals”.

3. Ali, Md. E.; Singharoy, A.; Datta, S. N. “Theoretical Investigation of Magnetic Properties of
Trimethylenemethane-Type Nitroxide Diradicals”.

4. Ali, Md. E.; Datta, S. N. “Theoretical Investigation of Magnetic Properties of Photomagnetic Molecules”.
Acknowledgment

I express my deep sense of gratitude, respect and admiration to my guide, Prof. S. N.


Datta, for his constant support and guidance, and for encouraging me to work independently
whenever it was required. I am greatly indebted to him for his priceless help, genial
behavior, moral boost, good wishes, kindness, ebullience, criticism, enormous patience,
useful suggestions and keen interests throughout the continuation of my work. His systematic
approaches to scientific problems and affirmative outlook have always enlivened me. The
affection, love and friendliness of ‘Kakima’ and Gargi deserve humbling admiration.
I also convey deepest gratitude and respect to Prof. Y. U. Sasidhar and Prof. Anindya
Datta for painstakingly evaluating my annual progress reports and bestowing their valuable
suggestions and comments throughout the PhD program.
I am indebted to Prof. Illas of University of Barcelona and Dr. Ciofini of CNRS Paris.
I learned certain scientific techniques from them. This work would have remained largely
incomplete without their help.
Susmit Basu has been associated in my work and is a considerate friend. I would also
like to thank my co-workers Anirban Panda, Nital Mehta, Ekram Hossain, Abhishek
Singharoy, Aritro Sinha Roy, Suvrajit Sengupta, Shubham Vyas, Sameer Patwardhan, Praket
Jha, Anshu Pandey, Prasuan Mukherjee and Dr. Anirban Misra.
I am eternally grateful to my “Abba” and “Maa” (parents), who in spite of great
hardships, have always encouraged me to pursue higher education and finally the PhD work.
Their blessings have caused the real motivation for me to join the program at IIT-Bombay. I
dedicate this thesis to them. I am also indebted to Amina, Rahmat and Barjahan for
continuous moral support. Little “Atiya” has been a constant source of joy. Last, but not the
least of all, Anjum Ismail has been a constant source of inspiration.

Md. Ehesan Ali


January 25th, 2007

Potrebbero piacerti anche