You are on page 1of 27

November 9, 2013 9:31 9in x 6in Theoretical and Computational Aspects. . .

b1639-ch02
Theoretical and Computational Aspects of Magnetic Organic Molecules Downloaded from www.worldscientific.com

CHAPTER 2

Organic Molecules, Radicals, and Spin States


by UNIVERSITY OF QUEENSLAND on 10/14/14. For personal use only.

Organic materials are of tremendous interest in a wide range of industrial


applications including the production of essential everyday items such as paper,
fuel, and adhesives, as well as the advancement of research and development in
plastics, pharmaceuticals, and biotechnology.
Most organic molecules and the materials they make up are diamagnetic
insulators, in keeping with their closed-shell electronic structure. It may seem
unlikely that such systems could have useful conductive, optical, or magnetic
properties similar to those of inorganic systems which make up so many modern
devices. In fact many organic materials, both naturally occurring and synthetic,
are found to have interesting and significant electrical, mechanical, and optical
properties. These characteristics make it possible to construct new devices with
potentially considerable advantages over existing products, including their light
weight, flexibility, and low cost. Some organic polymers, as flexible as any fiber,
are as conductive as copper. The combination of conductivity and flexibility has led
to the use of these polymers in the fabrication of electronic devices such as light-
emitting diodes and organic semiconductors [1]. Some organic charge transfer
salts are superconducting, with transition temperatures as high as 15 K [2], and are
therefore suitable for interesting applications.
One may still wonder whether organic materials can display useful levels of
magnetism. The conventional theory of ferromagnetism through spin exchange
applies most directly when ions have valence electrons in d- and f-shells. For
this reason, ferromagnetism is most often found in transition elements and
lanthanides. Traditional magnetic materials achieve ferromagnetism by the so-
called s–d interaction between conduction carriers and magnetic moments. It
would then appear that ferromagnetism in materials containing atoms of only
s- and p-valence electrons would be impossible. Indeed, elementary carbon does

15
November 9, 2013 9:31 9in x 6in Theoretical and Computational Aspects. . . b1639-ch02

16 Theoretical and Computational Aspects of Magnetic Organic Molecules

not have a spontaneous magnetic moment in any of its allotropic forms.1 Though
organic radicals have unpaired spins, they are known to be reactive. Extremely few
were found to be stable enough to form a crystalline solid. In the rare case of solid
formation, the intermolecular overlap of the singly-occupied molecular orbitals
(SOMOs) generally leads to dimerization, and to the formation of diamagnetic
Theoretical and Computational Aspects of Magnetic Organic Molecules Downloaded from www.worldscientific.com

molecular (weakly bound) solids with small negative Weiss constants. Therefore,
synthesis of organic ferromagnets was thought to be an impossible task.
In this chapter we will demonstrate that the pessimistic outlook is unjustified,
and that in fact there are real possibilities that magnetic materials of practical
value can be entirely organic. We first confront the obstacles to the development of
by UNIVERSITY OF QUEENSLAND on 10/14/14. For personal use only.

such magnets: the high reactivity of spin-bearing organic radicals, and the relatively
weak coupling of such species in condensed phases. Then we describe ways that the
obstacles have been overcome, eventually to produce room-temperature magnetic
materials.

Organic Monoradicals
Single organic molecules with S > 0 are spin carriers, and are called radicals.
Radical chemistry has been quite well known for close to a century. The first known
radicals were very reactive monoradicals, carrying a single S = 1/2 spin. Free rad-
icals can be produced by several methods including thermal excitation, photolysis
either with electromagnetic radiation or with particle radiations, photosensitization,
radiolysis, and by electrical discharges in tubes. They are usually detected by
rapid spectroscopic methods such as absorption spectroscopies, flash photolysis,
emissions, and electron spin resonance (ESR) or electron paramagnetic resonance
(EPR). Free radicals can also be detected and their properties can be described
by mass spectrometry, chemical methods, magnetic moment measurement, EPR
spectroscopy, and by trapping on a solid surface.
Examples of complex reactions that involve one or more free radical
intermediate are chain reactions like

H2 + Br 2 → 2HBr,
and (2.1)
H2 + Cl2 → 2HCl,

1 Recent discoveries of exotic organic materials have given a great boost to the study of organic
substances in the condensed phase. For example, when suitable dopants are added to fullerene (C60 ),
an allotrope of carbon, the material can become superconducting [3–5] or acquire an odd magnetic
behavior [1]. This is discussed in more detail below.
November 9, 2013 9:31 9in x 6in Theoretical and Computational Aspects. . . b1639-ch02

Organic Molecules, Radicals, and Spin States 17

as well as catalyzed chain reactions like the decomposition of acetaldehyde


catalyzed by iodine, the decomposition of hydrocarbon catalyzed by nitric oxide,
the decomposition of hydrogen peroxide catalyzed by ferrous ion, and the
decomposition of ozone catalyzed by chlorine [6].
Olefins normally polymerize by a free radical mechanism. The initiation may
Theoretical and Computational Aspects of Magnetic Organic Molecules Downloaded from www.worldscientific.com

be a second-order process involving two monomers or a monomer and a catalyst.


The propagation step involves the addition of a free radical to a double bond,
thereby creating another free radical:

R• + R CH=CH2 → •CHR =CH2 R. (2.2)


by UNIVERSITY OF QUEENSLAND on 10/14/14. For personal use only.

Finally, two radicals, different or identical, may combine to terminate the chain.
Thus the free radical (monoradical) chemistry is rich in information.
Despite their generally high reactivity, few persistent monoradicals are known.
The eldest is Gomberg’s triphenylmethyl radical [7]. Its persistence is due
to steric crowding, as is the lowered reactivity of the nitroxide radical in the
setting TEMPO ((2,2,6,6-tetramethylpiperidin-1-yl)oxyl) [8]. (See Figure 2.1.)
Stable radicals are of wide and increasing interest since they must be the key to
design and fabrication of molecular magnets [9].Although most magnetic materials
incorporate metals, purely organic systems are recognized. Thiazyl radicals [10–
11], di-tert-alkyliminoxyls [12], and delocalized radicals containing the hydrazyl
[R2 N-NR] unit [13] are of special interest across materials science and chemistry.

Gomberg’s triphenylmethyl (2,2,6,6-tetramethylpiperidin-l-yl)oxyl

di-tert-alkyliminoxyl hydrazyl
R

N R
R' NR2
O N

Figure 2.1. Some long-lived monoradicals.


November 9, 2013 9:31 9in x 6in Theoretical and Computational Aspects. . . b1639-ch02

18 Theoretical and Computational Aspects of Magnetic Organic Molecules

Diradicals
It is quite possible that a molecule may have more than one unpaired electron
associated with the nonbonding orbitals of different atoms. When these electrons
strongly interact with each other, they form a delocalized system of low spin. If the
Theoretical and Computational Aspects of Magnetic Organic Molecules Downloaded from www.worldscientific.com

interaction is weak, the unpaired electrons retain a great part of their nonbonding
nature, and give rise to diradicals and polyradicals. Of late, diradicals have become
known as common intermediates in chemical processes [14]. Polyradicals were
recognized more recently [15] and are becoming increasingly familiar [16].
A large number of organic diradicals are known, and many more can be
suggested. The biradicals can have singlet and triplet states. (The words biradical
by UNIVERSITY OF QUEENSLAND on 10/14/14. For personal use only.

and diradical are to be used synonymously throughout this book.) A few well-
known examples are given in Figure 2.2. The simplest one is methylene, which
has a triplet ground state; it serves as an illustration of Hund’s rule of maximum
spin multiplicity, because both the nonbonding orbitals in this species lie on
the same atom, that is, carbon. The singlet–triplet energy difference is about
8.3 kcal mol−1 [18]. Figure 2.3 shows some other well-known biradicals with
radical centers (X• and Y•) like •CH2 , •O, NO•, etc. Again, two of the same or
two different radical centers like the NN (nitronyl nitroxide), IN (imino-nitroxide),
VER (verdazyl), o-VER (oxo-verdazyl), and TTF (tetrathiafulvalene) moieties can
be joined together through a spacer or coupler that is a fragment of the olefins,
phenylene group, condensed aromatic units, and biphenyl etc. conjugated systems
so as to form a variety of diradicals.
These biradicals illustrate many of the fundamental structural units from which
very high-spin organic molecules systems are constructed. Whether the biradical
has a triplet ground state or not is key to the effective design of larger magnetic
systems. In the following sections we will describe some qualitative approaches to
the issue, and then turn to a more quantitative formulation based on the Heisenberg
Hamiltonian.

Intramolecular Interaction and Predicting the Spin State of Diradicals


The simplest of helpful guides for use in the prediction of spin preference depend
on the topology of a pi-system. Longuet-Higgins [19] explained the preference
in certain diradicals for a high-spin ground state by counting the number of
nonbonding molecular orbitals (NBMOs). His value is given by

N(NBMO) = NC − 2Ndb , (2.3)

where NC is the total number of conjugated carbon atoms and Ndb is the maximum
number of possible double bonds. Using Hund’s rule to assign electrons to
November 9, 2013 9:31 9in x 6in Theoretical and Computational Aspects. . . b1639-ch02

Organic Molecules, Radicals, and Spin States 19

TMM TME
Theoretical and Computational Aspects of Magnetic Organic Molecules Downloaded from www.worldscientific.com

X X X

Y
by UNIVERSITY OF QUEENSLAND on 10/14/14. For personal use only.

Ortho- Meta- Para-


Phenylene diradical

X Y X

Y
Symmetric Asymmetric
Biphenyl diradical

X Y X X

X Y Y

Y
Symmetric Asymmetric
Naphthalene diradical

J = 455 cm-1

Schlenk diradical

Figure 2.2. Well-known diradicals. TMM = tetramethylene methane, TME = tetramethylene ethane.
The coupling constant for the Schlenk diradical has been taken from Rajca [17].
November 9, 2013 9:31 9in x 6in Theoretical and Computational Aspects. . . b1639-ch02

20 Theoretical and Computational Aspects of Magnetic Organic Molecules

X X Y X

Y Y
Theoretical and Computational Aspects of Magnetic Organic Molecules Downloaded from www.worldscientific.com

S=0 S=1
(a)

* * * * *
X Y X

*
*
by UNIVERSITY OF QUEENSLAND on 10/14/14. For personal use only.

* *
Y
S=0 S=1
(b)

X Y X

S=0 S=1
(c)
Figure 2.3. Illustration of (a) the Longuet-Higgins rule, (b) the Ovchinnikov prescription, and (c) spin
alternation in unrestricted treatments.

the NBMOs, one can link the number of NBMOs to the maximum number
of unpaired spins. This is illustrated by Figure 2.3(a), where one can observe
that the paraphenylene structure has N(NBMO) = 0 while for metaphenylene,
N(NBMO) = 2. Consequently we assign paraphenylene a singlet ground state (no
unpaired electrons) and metaphenylene a triplet ground state. This simple but
elegant rule works well unless the ground state is an open-shell singlet.
The next major explanation came from the valence bond analysis of Ovchin-
nikov [20] who suggested a marking of the sites in a conjugated system as
alternately starred and unstarred carbon atoms such that the numbers of these
two groups of atoms satisfy n∗ > n. The spin number S can be determined by
S = (n∗ − n)/2, the excess spin of one type, up or down (see Figure 2.3(b)). This
is, of course, a useful analysis, but it is perhaps limited by the presumption that
the bonding is covalent. Besides, the valence bond approach is not as general or
theoretically systematic as the molecular orbital-based methodologies.
Borden and Davidson [21] brought the best calculations of the day to bear
on the prediction of spin state preference, and, by their recognition of disjoint
November 9, 2013 9:31 9in x 6in Theoretical and Computational Aspects. . . b1639-ch02

Organic Molecules, Radicals, and Spin States 21

and non-disjoint NBMOs, they resolved some cases for which preferences were
puzzling, notably the case of square cyclobutadiene.
Dougherty took a fresh view of the alternant hydrocarbon symmetries, and
suggested that in such pi-systems the atoms can be divided into two sets. Biradicals
can be formed by assigning the unpaired electrons only on the alternate (starred)
Theoretical and Computational Aspects of Magnetic Organic Molecules Downloaded from www.worldscientific.com

atoms such that they do not directly form a covalent bond [22]. In other words,
spin density resides only on alternate atoms. The biradical may or may not have
a triplet ground state. This reasoning overlooks spin polarization. It also gives no
indication of what would happen when the unpaired electrons are an odd number
of atoms apart.
by UNIVERSITY OF QUEENSLAND on 10/14/14. For personal use only.

Klein and March [23] investigated molecular magnetism via a valence bond
approach allowing unpaired electrons in the π- or σ-parts of an alternant network.
These unpaired electrons give qualitative information on the possibility of low-
lying high-spin states. The systems they have studied are conjugated radicals and
selected transition metal complexes. Klein and Alexander [24] proved new and
rigorous theorems for alternant systems. Alexander and Klein [24] considered π-
network hydrocarbons with dangling σ-bonds at carbene centers. These are again
based on the valence bond models of Pauling and Wheland [25]. These authors
made both qualitative and quantitative predictions of spin states for high-spin
carbenes.

An Alternative View of Spin State Preference


Perhaps the simplest rule is based on the observation that, in an unrestricted self-
consistent-field calculation of an organic system, the sign of the calculated spin
density alternates from one atom to the next [26]. This is illustrated by Figure 2.3(c).
In a time-dependent picture, we can consider spin waves propagating from
each of the two radical centers.A superposition of these waves appears as a standing
wave, the alternating alpha–beta pattern derived from unrestricted calculations.
The preferred spin state allows the two waves to interfere constructively. If the two
waves interfere destructively, forcing like spin at adjacent centers or a node in the
resultant superposition, then the spin assignment is disfavored.
Consider the case of substituted phenylene diradicals shown in Figure 2.3(c).
The alternating spin sequence reveals the intramolecular coupling to be antiferro-
magnetic for the para-isomer, and ferromagnetic for the meta-isomer.
Calculations of spin density in a wide range of systems [27] show that the spin
wave decreases in amplitude with increasing distance between the radical centers.
In fact, the amplitude is approximately inversely proportional to the distance, and
the spin density is proportional to the square of amplitude. This gives rise to a rapid
decrease of the coupling constant as the spacer size increases.
November 9, 2013 9:31 9in x 6in Theoretical and Computational Aspects. . . b1639-ch02

22 Theoretical and Computational Aspects of Magnetic Organic Molecules

The spin alternation is dampened when coupling is weakened in other ways,


as, for example, when the system is distorted so that conjugation is impeded. This
can occur when the radical carrier is twisted out of coplanarity to relieve steric
congestion.
If the substituents are bulky enough, as in the tert-butyl nitronyl nitroxides
Theoretical and Computational Aspects of Magnetic Organic Molecules Downloaded from www.worldscientific.com

described below, the NO radical site can be rotated so severely that conjugation
with the coupler will be cut off. This will prevent delocalization of the spin into the
ring and weaken the spin wave. A smaller spin coupling (smaller Jex value) results.
When the dihedral angles are quite large, the McConnell mechanism (described
below) imposes an antiferromagnetic nature on a stacking interaction. This is
by UNIVERSITY OF QUEENSLAND on 10/14/14. For personal use only.

clearly evident in orthophenylene nitronyl nitroxide diradicals [27, 28], and has
also been evident in a handful of metaphenylene diradicals [28, 29].
The spin alternation rule in unrestricted formalism is discussed in detail in
Chapter 5.

Polyradicals
Some triradicals and tetraradicals are shown in Figure 2.4, and polyradicals are
illustrated in Figure 2.5.
In general, for whatever is the case for diradicals, similar if not exactly identical
behavior is to be found with polyradicals. Thus the topological effect, influence of

E(D)-E(Q)=3J E(D)-E(Q)=J
Triradicals

O O

Tetraradicals
Figure 2.4. Triradicals and tetraradicals. The first tetraradical shown here is a radical derivative of
triphenylmethyl, the second is an extended form of metaphenylene and the third is an example of a
fused ring polyradical. Quartet states (S = 3/2) and quintet states (S = 2) are available to these systems.
November 9, 2013 9:31 9in x 6in Theoretical and Computational Aspects. . . b1639-ch02

Organic Molecules, Radicals, and Spin States 23

Ar Ar Ar Ar Ar

Ar Ar
Theoretical and Computational Aspects of Magnetic Organic Molecules Downloaded from www.worldscientific.com

Ar Ar Ar Ar

Ar Ar

Ar Ar Ar Ar Ar
by UNIVERSITY OF QUEENSLAND on 10/14/14. For personal use only.

S=6
Polyradicals

Figure 2.5. Significant polyradicals. States with S = 6 and higher are available [16, 17].

N N N N N

O O O O O

N N N N N

O O O O O

Figure 2.6. Nitroxide chains. When — represents a tert-butyl group, J > 300 cm−1 for the biradical,
and J ∼ 170 cm−1 for the triradical [29].

the spacer size, and the low-spin–high-spin energy difference follow similar trends
in diradicals and polyradicals.
It is interesting to note the decrease in the coupling constant attending
the extension of a polyradical chain, illustrated by the metaphenylene-coupled
nitroxide chains shown in Figure 2.6 [29]. The decrease may be due to departures
from planarity as the chain grows. This has ominous implications for the feasibility
of constructing systems with high spin density and strong coupling. This problem
is explored in Chapters 6, 7, 8, and 11.
We now turn to a more quantitative characterization of diradicals.
November 9, 2013 9:31 9in x 6in Theoretical and Computational Aspects. . . b1639-ch02

24 Theoretical and Computational Aspects of Magnetic Organic Molecules

Hamiltonian for Organic Diradicals


The effective spin Hamiltonian for an organic biradical in a magnetic field is the
sum of two terms: the Heisenberg exchange Hamiltonian which represents the
interaction of the two unpaired spins, and the paramagnetic contribution from
Theoretical and Computational Aspects of Magnetic Organic Molecules Downloaded from www.worldscientific.com

interaction of the total spin angular momentum with an external magnetic field:

H spin = H ex + H para ,
H ex = E0 − 2Jex S 1 .S 2 , (2.4)
H para = −µS .B a .
by UNIVERSITY OF QUEENSLAND on 10/14/14. For personal use only.

Here, Jex is called the intramolecular magnetic exchange coupling constant, and
it should not be equated with the inter-species coupling constant discussed in
Chapter 1. This is often a major source of confusion. We find Jex = 1/2[E(S =
0) − E(S = 1)] and E0 = 1/4[E(0) (S = 0) + 3E(0) (S = 1)], where E(0) (S = 0)
and E(0) (S = 1) are the energy values in the absence of a field. When the magnetic
field is switched on, we get the energy levels:

3
E(S = 0, MS = 0) = E0 + Jex
2
1
E(S = 1, MS = 1) = E0 − Jex − µ1 Ba
2
1 (2.5)
E(S = 1, MS = 0) = E0 − Jex
2
1
E(S = 1, MS = −1) = E0 − Jex + µ1 Ba ,
2

where µ1 = ge βe . The coupling is said to be “ferromagnetic” when Jex is positive.


In that case, the electronic ground state has parallel spins. A negative Jex gives an
“antiferromagnetic” coupling with a ground state having spins antiparallel to each
other.
For an intramolecular ferromagnetic (FM: parallel spin) coupling (Jex > 0),
the energy levels are shown in Figure 2.7. Changing the sign of Jex inverts the
diagram.
A competition between the exchange interaction in Eq. (2.4) and thermal
agitation determines the magnetic moment. Ensembles of organic radicals coupled
either ferromagnetically or antiferromagnetically will show paramagnetic behavior
at high temperature such that |Jex |/τ << 1; then the ordering due to the coupling
is overwhelmed by thermal disorder. This is the weak field limit.
November 9, 2013 9:31 9in x 6in Theoretical and Computational Aspects. . . b1639-ch02

Organic Molecules, Radicals, and Spin States 25

S=0, MS=0

3J/2
Theoretical and Computational Aspects of Magnetic Organic Molecules Downloaded from www.worldscientific.com

E0
J/2 S=1, MS=1
g|β|B
S=1, MS=0
g|β|B
S=1, MS= -1
by UNIVERSITY OF QUEENSLAND on 10/14/14. For personal use only.

Figure 2.7. Splitting of the levels for a biradical with Jex > 0, in the absence and in presence of an
applied field.

Using the Boltzmann distribution for NS,MS /N, and the magnetization M as


1 
S
M=− µ1 MS NS,MS , (2.6)
S=0 MS =−S

for a weak field and a ground state triplet one obtains the expression

2Nµ21
χ=   , (2.7)
3+e−2Jex /τ τ

for the susceptibility that is dimensionless. This equation is known as the Bleaney–
Bowers equation [30]. The molar susceptibility becomes

3.007
χM =   , (2.8)
3+e−2.8775Jex /T T

where Jex is in cm−1 and T is in Kelvin. The unit of molar susceptibility is


cm3 mol−1 . For Jex > 0 (FM coupling), one finds 0.75/T ≤ χM ≤ 1/T whereas
for Jex < 0 (antiferromagnetic (AFM) coupling), one obtains 0.75/T ≥ χM ≥ 0.
The χM versus 1/T plot is shown in Figure 2.8.
The energetics and the consequent magnetization of polyradicals can be
similarly worked out, although the task would be more complex.
Our chief interest lies in molecules that are internally ferromagnetically
coupled. These can form either ferromagnetic or antiferromagnetic crystals, and
show paramagnetism in solution or in gel.
November 9, 2013 9:31 9in x 6in Theoretical and Computational Aspects. . . b1639-ch02

26 Theoretical and Computational Aspects of Magnetic Organic Molecules

χΜ (in cc mol-1)
Theoretical and Computational Aspects of Magnetic Organic Molecules Downloaded from www.worldscientific.com

Figure 2.8. Molar susceptibility of a paramagnetic solid of biradicals versus 1/T .


by UNIVERSITY OF QUEENSLAND on 10/14/14. For personal use only.

From Molecules to Materials


Single molecules can have high-spin states, but that does not make them
magnetic materials. Assembly into extended structures is required. We have
already mentioned why this seems impossible, since radicals can couple to form
conventional closed-shell systems. Even if molecules retain their spin identity
in the solid phase, one needs ferromagnetic coupling between the molecules for
macroscopic magnetism to emerge. Therefore, assembly of radicals into organic
paramagnets, let alone ferromagnets, would seem to be an impossible task.
However, what was deemed impossible by conventional wisdom was accomplished
in nature. Furthermore, solid-state design can apply the lessons learnt from nature
to go beyond what she has yielded.

Early Discoveries of Organic Ferromagnetic Systems


The species described below are organic molecular magnets, which is the
name given to a cluster or a crystal of weakly interacting molecules with a
large equilibrium population of high-spin states. It may behave as a permanent
paramagnet, especially when the molecules lack some features of symmetry. Often
it is a ferromagnet; that is, a magnetic moment is induced parallel with an applied
magnetic field. In the absence of a field, the magnetization is observed to relax
very slowly.
The organometallic substances [Mn4 O3 Cl4 (CH3 CH2 COO)3 (py)3 ]2 (abbrevi-
ated [Mn4 ]2 , S = 9/2), a mixed valence manganese-oxo cluster Mn12 O12 (CH3
COO)16 (H2 O)4 , (abbreviated Mn12 ac, S = 10) [31], and an iron(III) oxo-
hydroxo cluster with the macrocyclic ligand tacn, [(tacn)6 Fe8 O2 (OH)12 (H2 O)]8+ ,
(abbreviated Fe8 , S = 10) [32] are well-known single molecule magnets. Among
November 9, 2013 9:31 9in x 6in Theoretical and Computational Aspects. . . b1639-ch02

Organic Molecules, Radicals, and Spin States 27

O O
Theoretical and Computational Aspects of Magnetic Organic Molecules Downloaded from www.worldscientific.com

N N

O O

Fullerene Galvinoxyl NITR


by UNIVERSITY OF QUEENSLAND on 10/14/14. For personal use only.

NO2 R1
O
O N

O
O

N O
O O
O
N N N N

N N
R2 R2

p-NPNN Tanol suberate Verdazyl

Figure 2.9. Important molecules known to form magnetic materials.

purely organic substances, p-NPNN [2-(4-nitrophenyl)-4,4,5,5-tetramethyl-4,5-


dihydro-1H-imidazolyl-1-oxy-3-oxide] was one of the first organic ferromagnets
to be discovered [33–35]. In Figure 2.9 we show common molecules known to
form stable organic magnetic materials. Apart from fullerene, which is a closed-
shell system with no net spin, the five molecules in Figure 2.9 are long-lived
radicals: NITR (α-nitronyl nitroxide), p-NPNN, and verdazyl are monoradicals
(with one unpaired electron in each species), while galvinoxyl and tanol suberate
are biradicals (two unpaired electrons).
In 1969, solid galvinoxyl was found to have a positive Weiss constant of
about 19 K [36]. The intermolecular interaction is manifestly ferromagnetic.
Galvinoxyl crystal has the nearly planar diradicals stacked along the c-axis.
A ferromagnetic Heisenberg chain with J = 13 ± 1 kB K can easily explain the
magnetic susceptibility at high temperature [37]. (It is worthwhile to note here
1 kB K = 0.695 cm−1 .) A first-order phase transition in the solid occurs at 85 K. The
low-temperature phase is characterized by a strong antiferromagnetic interaction
with J ≈ −260 kB K [38, 39]. These observations can be explained in terms of
a thermal equilibrium between singlet and triplet states of the diradical, singlet
being more stable [37].
November 9, 2013 9:31 9in x 6in Theoretical and Computational Aspects. . . b1639-ch02

28 Theoretical and Computational Aspects of Magnetic Organic Molecules

N
N
Theoretical and Computational Aspects of Magnetic Organic Molecules Downloaded from www.worldscientific.com

O O O O O O
N N N N N N

Figure 2.10. Structures of a few important magnetic molecules based on NN, other than NITR and
by UNIVERSITY OF QUEENSLAND on 10/14/14. For personal use only.

p-NPNN.

Nitronyl nitroxides (NN), shown as NITR in Figure 2.9, led to the preparation
of real organic ferromagnets [40]. Figure 2.10 shows a few more variations. These
radicals are stable, and indeed have found application as spin labels in biological
systems. The MO with the odd electron — the SOMO — is mainly spread over the
NO moities and has a node on the carbon atom between them. Nitronyl nitroxide
solids generally do not have long-range ferromagnetic order. This is why the
discovery of long-range ferromagnetism in the β phase of the crystal of para-
nitrophenyl NN (p-NPNN) caused a great sensation. It was viewed as a giant
step in the investigation of organic materials, albeit at a very low ferromagnetic
transition temperature (0.65 K) [34]. In fact, the transition temperatures of all NN-
based monoradical solids with long-range magnetic order have been found to be
extremely low.
Tanol suberate is another important biradical long known in this field, initially
viewed as a promising candidate. The spin density is almost equally shared between
the two atoms of the NO group [41]. The Curie–Weiss law is obeyed by the crystal
of this molecule. The Curie temperature is 0.7 K. A λ anomaly in the specific heat
was observed at 0.38 K by Saint-Paul and Veyret in 1973 [42]. This phenomenon
was attributed to a ferromagnetic spin alignment. In reality, tanol suberate is an
antiferromagnet. It undergoes a metamagnetic transition (to ferromagnetism) in a
field of 6 mT [43, 44]. Hence it is mainly of theoretical interest.
Verdazyls (VER) form another group of organic radicals that are quite stable.
In 1990, it was discovered that solid (1-nitrophenyl)3,4,5-triphenylverdazyl is
ferromagnetic with Weiss constant 1.6 K whereas (1-nitrophenyl)3,5-dipheny-
lverdazyl is antiferromagnetic [45]. This is an example of molecular engineering,
that is, crystal packing with a lateral shift to promote ferromagnetic interaction
between nearest neighbors. Other related species like o-VER and thioxo-verdazyl
are well known [46]. A variety of magnetic systems such as ferromagnets,
November 9, 2013 9:31 9in x 6in Theoretical and Computational Aspects. . . b1639-ch02

Organic Molecules, Radicals, and Spin States 29

antiferromagnets, and materials with spin-Peierls transitions can be prepared from


substitutions at different positions of VER and its oxo- and thioxo- variants. Two
examples of this are:

1. The compounds p-CDpOV, p-CDTV and p-MeDpOV are ferromagnetic with


Theoretical and Computational Aspects of Magnetic Organic Molecules Downloaded from www.worldscientific.com

TC 0.21 K, 0.68 K and 0.67 K, respectively [47, 48].


2. Alloys of substituted o-VER radicals have been used to investigate the effect
of a random distribution of ferromagnetic and antiferromagnetic interactions in
an effective one-dimensional chain [49].
by UNIVERSITY OF QUEENSLAND on 10/14/14. For personal use only.

Radicals containing sulfur are increasingly emerging as good candidates for


molecular magnets. Spontaneous magnetization below 35 K has been found in
dithia-diazolyl radical p-NC(C6 F4 ) (CNSSN) [50]. This is a far cry from the small
TC solids of nitroxides and VERs, yet it is by no means the whole story. The
transition temperature increases to 65 K under a pressure of 16 kbar [51]. The
radical cation tetrathiafulvalene (TTF) has been attracting attention from both
experimental and theoretical chemists [2, 52]. This is discussed in Chapters 6, 7,
and 8.
The reader may wonder why fullerene is included in this collection of systems
proven to be ferromagnetic solids. After all, fullerene has a closed-shell electronic
structure and in pure form is a diamagnetic nonconductor. The story is one worth
telling.
In 2001, Makarova et al. [53] reported room-temperature ferromagnetism in
C60 fullerene that had been polymerized at a high temperature and pressure. This
discovery was subsequently verified by Wood et al. [54] and Narozhnyi et al. [55].
Buckyballs collapse a little above 800 K. Samples polymerized a little below
this temperature have layers of covalently bonded C60 molecules forming a two-
dimensional rhombohedral polymer phase, and show a maximum in magnetization
that is nevertheless small and nonuniform. The Curie temperature is found to range
from 500 K [53] to 820 K [54]. This high TC indicates that the coupling constant is
quite strong. Electron microscopy and X-ray diffraction experiments show that the
buckyballs remain largely intact in the magnetic phase [54]. These observations can
be explained as follows. First, the polymer has some of the intermolecular bridging
bonds broken. These so-called “dangling bonds” serve as radical centers. The
random distribution of the radical centers is responsible for the lack of uniformity of
magnetization. Second, the low density of the dangling bonds and the large volume
of the buckyballs make the magnetization very small [52–55]. Nevertheless, there
exists in the polymer a strong conjugation mediated by the buckyballs. We will see
in Chapters 5, 6, 7, and 8 that a strong and extensive conjugation leads to a strong
November 9, 2013 9:31 9in x 6in Theoretical and Computational Aspects. . . b1639-ch02

30 Theoretical and Computational Aspects of Magnetic Organic Molecules

magnetic coupling that is associated with a high Curie temperature. This tempts
us to imagine using the buckyball as a coupler.
Now we know about some solids that are ferromagnetic — two questions
arise: how does it happen, and how can we make it happen?
Theoretical and Computational Aspects of Magnetic Organic Molecules Downloaded from www.worldscientific.com

Intermolecular Interaction and Ferromagnetism in Extended Systems


Assuming a generalized exchange Hamiltonian, McConnell proposed a theo-
retical model for spin density–spin density interaction in a stack of π-electron
molecules [56]. He approximated the spin Hamiltonian for a pair of molecules A
by UNIVERSITY OF QUEENSLAND on 10/14/14. For personal use only.

and B as

H AB = −S A · S B JijAB ρiA ρjB , (2.9)
ij

where S A and S B are the total spin operators, and ρiA and ρjB are the π-spin densities
on atoms i and j of the respective molecules (shown as superscripts). Positive
and negative spin densities are found on different atoms of each molecule. This
is known as spin polarization. Also, the magnetic exchange interaction between
corresponding atoms, JiiAB , is generally negative for aromatic molecular stacks.
Therefore, when the molecules are exactly stacked on each other, the intermolecular
exchange interaction energy becomes negative when S A and S B are antiparallel,
that is, the molecules are antiferromagnetically coupled. When the stacking is
such that the atoms of positive spin density of one molecule are most strongly
coupled to atoms of negative spin density in the neighbor, the intermolecular
interaction energy becomes negative for parallel spins. The stacked molecules
are then ferromagnetically coupled to each other.
Indeed from a detailed analysis it was found that a ferromagnetically coupled
stacking requires (i) conjugated atoms on each molecule, (ii) capability of the
nonbonding π orbital to be coextensive when another molecule approaches from
a lateral direction, and (iii) positive exchange integrals between the most strongly
coupled pairs of atoms [57]. These are reminiscent of the conditions presented by
Dougherty [22].
Another way of viewing intermolecular ferromagnetic interaction was sug-
gested by Awaga, Sugano, and Kinoshita (ASK) [58]. As with McConnell’s second
suggestion [59], this view relies on the preferential stabilizing influence on a triplet
state by charge transfer. To show how this can happen in a stack of planar radicals
such as galvinoxyl, first consider a pair G and G as two doublets, each described
by restricted open-shell Hartree–Fock (ROHF) wave functions. Then at each site
there is a SOMO (SO) with one alpha electron. Other electrons occupy MOs with
November 9, 2013 9:31 9in x 6in Theoretical and Computational Aspects. . . b1639-ch02

Organic Molecules, Radicals, and Spin States 31

common spatial factors in alpha–beta pairs. The highest-energy MO in this pair set
we call NHOMO (NH) and the lowest-energy entirely vacant MO we call NLUMO
(NL), following ASK’s system of labeling.
When the ROHF restriction is lifted, the unrestricted Hartree–Fock (UHF)
alpha and beta MOs differ in energy and spatial form. Then in the alpha manifold
Theoretical and Computational Aspects of Magnetic Organic Molecules Downloaded from www.worldscientific.com

the NLUMO-α is vacant and both SOMO-α and NHOMO-α are singly occupied.
Their counterparts in the beta manifold differ in form and energy; only NHOMO-β
is occupied, while SOMO-β and NLUMO-β are vacant.
ASK consider a pair of doublet radicals G and G with spins oriented
either parallel or antiparallel, triplet or singlet. The states are called NT and NS
by UNIVERSITY OF QUEENSLAND on 10/14/14. For personal use only.

respectively, the letter ‘N’ representing ‘normal’. The arrangement of electrons


could be specified as in Figure 2.11. Several kinds of charge-transfer (CT)
excitations can be envisioned:

1. From NS, an electron can pass from SO-α of G to SO-α of G . Both cation G and
anion G are then singlets. This excitation produces singlet S0 and the CT can
mix with and enhance the stability of NS. It is an antiferromagnetic influence.
2. From NT, an electron can pass from NH-β of G to SO-β of G . Cation G is a
triplet while anion G is a singlet. The composite system is a triplet called T1
and the CT excitation stabilizes NT. It is a ferromagnetic influence.
3. From NS, an electron can pass from NH-α of G to SO-α of G . Cation G and
anion G are both singlets, the composite system is a singlet called S1 and the
CT excitation stabilizes NS. It is an antiferromagnetic influence.
4. From NT, an electron passes from SO-α of G to NL-α of G . Cation G is a
singlet and anion G is a triplet. The composite system is a triplet called T2 and
the CT excitation stabilizes NT. It is a ferromagnetic influence.
5. From NS, an electron passes from SO-α of G to NL-α of G . Cation G and
anion G are both singlets, the composite system is a singlet called S2 and the
CT excitation stabilizes NS. It is an antiferromagnetic influence.

Which of these competing influences will dominate depends on the ease of each
specific charge transfer. In the ROHF limit there will be good energy matching of
SOMO-α on G with SOMO-α on G . Under those circumstances antiferromagnetic
coupling is established and the singlet state is preferentially stabilized by CT type 1.
In galvinoxyl, ASK found that strong spin polarization due to a strong
intramolecular exchange shifted MO energies such that the SOMO-α lies close to
(or even slightly below) the NHOMO-β. In other words NH-β of G and SO-β of
G were very well energy-matched. This would favor CT of type 2 or 3. Owing
to the lower energy of the triplet T1 compared with S1, we expect preferential
November 9, 2013 9:31 9in x 6in Theoretical and Computational Aspects. . . b1639-ch02

32 Theoretical and Computational Aspects of Magnetic Organic Molecules


Theoretical and Computational Aspects of Magnetic Organic Molecules Downloaded from www.worldscientific.com
by UNIVERSITY OF QUEENSLAND on 10/14/14. For personal use only.

α β α β α β β α
Figure 2.11. Radical dimer coupled by charge transfer interactions. This diagram illustrates the
scheme of Awaga, Sugano, and Kinoshita [58].
November 9, 2013 9:31 9in x 6in Theoretical and Computational Aspects. . . b1639-ch02

Organic Molecules, Radicals, and Spin States 33

stabilization of NT, unless CT of type 1 dominates. The ferromagnetism of solid


galvinoxyl owes its origin not only to favorable energy matching making CT of
type 2 very effective, but also to the very small overlap between the SOMOs
of the nearest neighbors. This weakens the effects of charge transfer of type 1,
predominant in many crystals of radicals.
Theoretical and Computational Aspects of Magnetic Organic Molecules Downloaded from www.worldscientific.com

Although ferromagnetic interactions in crystals of organic radical are still


rarities, Kinoshita used these ideas to rationalize the ferromagnetism of solid
p-phenyl nitronyl nitroxide [60].

Molecular Design and Molecular Magnets


by UNIVERSITY OF QUEENSLAND on 10/14/14. For personal use only.

The whole field changed with the advent of molecular design. Indeed, molecule-
based magnets [61] have become quite well known at present. Molecular
engineering plays a crucial role in the synthesis of these materials. Its first
challenge is to deal with the intrinsic reactivity of the radical. Neutral organic
radicals are highly reactive because of the unpaired electron(s). As already noted,
triphenylmethyl radical was one of the earliest organic radicals prepared [7], and a
number of diradical and polyradical derivatives have been known [16]. However,
they are not of much use because of their high reactivity. Organic radicals can be
made stable in two ways:

1. By the addition of aromatic rings which delocalizes the unpaired electron.


2. By the introduction of bulky substituents that bar the approach of reacting
species to the radical center.

These precautions can prevent the formation of dimers with zero net spin.
A major difference between magnetic molecules of organic and inorganic
origins has been mentioned in Chapter 1, namely that the source for magnetism
is S for organic molecules and J for inorganic complexes. We now mention
another important difference. Organic radicals in different spin states generally
differ in molecular geometry whereas transition metal complexes in different
angular momentum states have more or less the same geometry.2 Thus, an organic
molecular solid having more than one spin state coexisting in thermal equilibrium
often undergoes a phase transition on increasing or decreasing temperature. Such
phase changes are less common to inorganic materials.
Let us assume that these preliminary (although serious) difficulties were to be
overcome. Solids of neutral organic radicals are mostly found to be paramagnetic as

2 The geometry of the metal complex can change more appreciably when the oxidation state of the
metal atom changes.
November 9, 2013 9:31 9in x 6in Theoretical and Computational Aspects. . . b1639-ch02

34 Theoretical and Computational Aspects of Magnetic Organic Molecules

long as dimers are not formed. In some cases, these crystals exhibit ferromagnetic
or antiferromagnetic behavior.
Suppose that in a dimer composed of two monomers in parallel planes, each of
the two partners’magnetic sites are antiferromagnetically coupled.A ferromagnetic
coupling would result by means of a lateral slide of one monomer with respect to the
Theoretical and Computational Aspects of Magnetic Organic Molecules Downloaded from www.worldscientific.com

other. One may introduce a suitable functional group at one end of a six-membered
ring such that it sticks out of the ring plane on one side. Thus, one may exploit the
steric effect to get an angle shaped structure that is accompanied by the desired
lateral slide [45]. This can produce ferromagnetic coupling between successive
species in a stack. As additional applications of molecular engineering, one may
by UNIVERSITY OF QUEENSLAND on 10/14/14. For personal use only.

consider:

1. Adding specialized molecular fragments.


2. Making a small adjustment to the structure.
3. Synthesizing extended systems.
4. Using the characteristics of molecular crystals.

Like inorganic materials, organic molecular magnetic materials can have a


wide range of properties. They can be diamagnets, paramagnets, ferromagnets,
ferrimagnets, or antiferromagnets. Organic molecules can be used in another
front of molecular magnetism. As ligands, they can be varied in transition metal
complexes to modify the magnetic behavior. They can also be used to mediate the
magnetic interaction between two metal ions in the same complex, or even between
two different complexes.
Using suitable molecular fragments, one can synthesize optically active
magnets [62, 63], magnetic sensors, or conducting magnetic systems [64–66]. The
coexistence of charge carriers and magnetic moment gives rise to a possibility
of spintronics with immense scope in the computer industry since information
would be carried by a 4-bit process instead of the 2-bit one common in electronic
devices. On one hand, a small adjustment to the structure can tune the molecule to
a particular property. On the other hand, a desired material can be tailor made by a
set of structural alterations [61]. To top it all, one may make different networks like
polymer (one-dimensional), sheet (two-dimensional), crystal (three-dimensional),
or cluster (three-dimensional). This variety is needed not only for a comprehensive
understanding of magnetism in substances but also for a plethora of applications.
Systems with low dimensionality are considered necessary for investigation of spin
chains, spin-Peierls transition and spin ladders [67–69]. Last but not least, some
properties of molecular crystals can vary smoothly with pressure, owing to their
rather weak intermolecular forces. This seems analogous to the responsiveness of
November 9, 2013 9:31 9in x 6in Theoretical and Computational Aspects. . . b1639-ch02

Organic Molecules, Radicals, and Spin States 35

some tunable dye lasers. As a result, organic magnetic materials can have a wide
range of applications.
Excellent discussions on how one may employ organic radicals as building
blocks for functional materials can be found in a collection of essays edited
by Hicks [9] and a recent review by Ratera and Veciana [70]. The discussion
Theoretical and Computational Aspects of Magnetic Organic Molecules Downloaded from www.worldscientific.com

in this chapter has been specialized and limited, in view of our aims firstly to
identify the features of organic radicals that can help or hinder the assembly of
organic ferromagnetic materials and secondly to lay a foundation for more detailed
discussion of the electronic and magnetic states of these systems. Complementary
discussions of organic radicals and molecular magnetic states are given by
by UNIVERSITY OF QUEENSLAND on 10/14/14. For personal use only.

Iwamura [71], Buchachenko [72], and Iwamura and Koga [73].

Experimental Methods for Triplets


Although our preoccupation and central focus in our discussion is theory
and computation, it would be remiss if we did not acknowledge the striking
developments in experimental study of high-spin molecules that attended and
encouraged conceptual advances. We rely on such data to stay in contact with
factual reality.
An excellent place to start is with the survey of spectroscopic theory and
experimental study of the triplet state by McGlynn, Azumi, and Kinoshita [74].
Luminescence data, including phosphorescence polarization, and lifetimes, can
define the orbital makeup of the triplet state, diagnose the degree and impact of
spin–orbit coupling, and provide estimates of the singlet–triplet gap. Absorption
spectroscopy can give a complementary picture, since the selection rule barring
direct transitions from the ground state singlet to any triplet state, although strict,
is not absolute.
Since high-spin species are often very reactive, development of matrix
isolation techniques [75–78]3 and instruments for rapid measurements [79] was a
major boost to the study of radicals and polyradicals.
Characterization of states other than a singlet is the natural territory of electron
paramagnetic resonance spectroscopy. Simulation of observed spectra is aided by
a Heisenberg Hamiltonian. Analysis yields two significant parameters D and E;
these define the departure of magnetic sublevels from perfect degeneracy even in
the absence of an external field (hence the name, zero-field splitting). These are
still in routine use in the study of triplet systems.4

3 For applications to triplets see Bally [78].


4 For one example, among many, see Iwamoto, Hirai, and Tomioka [80].
November 9, 2013 9:31 9in x 6in Theoretical and Computational Aspects. . . b1639-ch02

36 Theoretical and Computational Aspects of Magnetic Organic Molecules

When the singlet–triplet gap is comparable with thermal energy, the EPR
signal intensity follows a temperature-dependence described by:
C 3exp(− EST /kB T)
I= . (2.10)
T [1 + 3exp(− EST /kB T)]
Theoretical and Computational Aspects of Magnetic Organic Molecules Downloaded from www.worldscientific.com

This is the Bleaney–Bowers equation discussed earlier in this chapter. At a


sufficiently high temperature ( EST /kB T << 1) it turns into the Curie law.
Extraction of the singlet–triplet gap from such data is often difficult in practice.
Cyclic voltammetry is a frequent adjunct to investigations of open-shell
molecules, allowing description of responses of systems to the injection or
by UNIVERSITY OF QUEENSLAND on 10/14/14. For personal use only.

withdrawal of single electrons [81].


Magnetochemistry [82] is an indispensible experimental method for the study
of radicals and derived materials [83]. The advent and improvements in SQUID
(superconducting quantum interference) magnetometers allow measurements of
high accuracy and sensitivity [84]. The methods widely used in study of high-
spin molecules were succinctly described by Iwamura, Koga, and Matsuda [85].
Direct measurement of magnetization M and magnetic susceptibilities allows
the inference of the spin quantum number, as from inspection of M versus
the ratio of magnetic field to temperature Bapp /T (see Eqs (1.22)–(1.25)). A
plot of susceptibility χ (or its inverse) against temperature T , or alternatively a
plot of the product of susceptibility and temperature χT against T will define
the system’s spin state. See the discussion under the Bleaney–Bowers form,
Eqs (2.6)–(2.8).
These methods have long since been widely used. A more recent and
revolutionary development in the experimental study of the singlet–triplet gap is
negative ion photoelectron spectroscopy (NIPES) [86]. The method gives directly
the relative energies of the neutral species derivable from the anion; if the anion
is a doublet, then both singlet and triplet are among the accessible state of
the neutral species generated by photodetachment. A detailed description of the
apparatus, along with a landmark application to the evaluation of the singlet–
triplet gap in methylene and its deuterated isotopomer was reported in 1985 [87].
Access to high quality data for gas phase species, attended by the development of
computational methods and theories capable of describing high-spin systems has
advanced understanding substantially [88].
The physics of the NIPES experiment is shown in Figure 2.12(a) for a singlet
anion subjected to light of sufficient energy to detach one electron. On the right,
energy is injected into an anionic system. The strongly energized system can lose
an electron and occupy a vibronic state of the neutral species. The energy hν of the
photon minus the maximum kinetic energy of the electron eKE defines the relative
November 9, 2013 9:31 9in x 6in Theoretical and Computational Aspects. . . b1639-ch02

Organic Molecules, Radicals, and Spin States 37

E = hν

eKE
eKE
Theoretical and Computational Aspects of Magnetic Organic Molecules Downloaded from www.worldscientific.com

R
eBE

Observed
Spectrum EA (R )
E
R
by UNIVERSITY OF QUEENSLAND on 10/14/14. For personal use only.

(a)
Photoelectron Counts

Electron Binding Energy, eV

(b)
Figure 2.12. (a) Schematic illustration of events in negative ion electron detachment spectroscopy.
Reproduced with permission from Wenthold and Lineberger [86]. Copyright 1999 American Chemical
Society. (b) Negative ion photodetachment spectrum for metaxylylyl anion, showing singlet and triplet
states of metaxylylene diradical. Adapted with permission from Wenthold, Kim, and Lineberger [89].
Copyright 1997 American Chemical Society.

energy of the anion and the neutral molecule, that is, the electron affinity. The eKE
spectrum will show vibrational structure.
Figure 2.12(b) shows the case of metaxylylyl anion [89], which upon
photodetachment of an electron forms the neutral with two accessible singlets as
well as the triplet ground state. The binding energy plotted on the horizontal axis
is the energy removed from the photon. The minimum electron binding (at right)
corresponds to the lowest-energy state, and the greatest energy transfer to the
emitted electron. The vibrational progression in that (most stable) state steps to the
November 9, 2013 9:31 9in x 6in Theoretical and Computational Aspects. . . b1639-ch02

38 Theoretical and Computational Aspects of Magnetic Organic Molecules

left, until the strong peak corresponding to a higher-energy state appears. In this
case the upper 1A1 state is about 16 kcal mol−1 higher than the 3 B2 ground state
triplet. The open-shell 1 B2 lies at about 35 kcal mol−1 .
Theoretical and Computational Aspects of Magnetic Organic Molecules Downloaded from www.worldscientific.com

References
[1] Blundell, S. J.; Pratt, F. L. J. Phys. Condens. Matter 2004, 16, R771–R828.
[2] Enoki, T.; Miyazaki, A. Chem. Rev. 2004, 104, 5449.
[3] Coronado, E.; Delhaes, P.; Gatteschi, D.; Miller, J. S. Eds., Molecular Magnetism:
From Molecular Assemblies to the Devices, Kluwer, Dordrecht, 1996; Miller, J.;
Gatteschi, D. Chem. Soc. Rev. 2011, 40, 3053.
by UNIVERSITY OF QUEENSLAND on 10/14/14. For personal use only.

[4] Ishiguro, T.; Yamaji, K.; Saito, G. Organic Superconductors, Springer, Berlin, 1998.
[5] Rosseinsky, M. J. J. Mater. Chem. 1995, 5, 1497.
[6] Laidler, K. J. Chemical Kinetics, Tata-McGraw-Hill, New Delhi, 1991.
[7] Gomberg, M. J. Am. Chem. Soc. 1900, 22, 757; Gomberg, M. J. Am. Chem. Soc. 1901,
23, 496; Gomberg, M. Chem. Rev. 1924, 1, 91; Gomberg, M. J. Am. Chem. Soc. 1902,
24, 597.
[8] Lebedev, O. L.; Kazarnovskii, S. N. Zhur. Obshch. Khim. 1960, 30(5), 1631.
[9] Hicks, R. Ed., Stable Radicals: Fundamentals and Applied Aspects of Odd-electron
Compounds, Wiley, New York, 2010.
[10] Rawson, J. M.; Palacio, F. Structure and Bonding 2001, 100, 93–128.
[11] Rawson, J. M.; Alberola, A.; Whalley, A. J. Mater. Chem. 2006, 16, 2560.
[12] Ingold, K. U. In Stable Radicals: Fundamentals and Applied Aspects of Odd-Electron
Compounds, Hicks, R. Ed., Wiley, New York, 2010, 231–244.
[13] Hicks, R. G. In Stable Radicals: Fundamentals and Applied Aspects of Odd-Electron
Compounds, Hicks, R. Ed., Wiley, New York, 2010, 317–380.
[14] Klessinger, M.; Michl, J. Excited States and Photochemistry of Organic Molecules,
Wiley-VCH, New York, 1994.
[15] Leo, M. Ber. Deutsche Chem. Ges. 1937, 70B, 1691.
[16] Rajca, A. Chem. Eur. J. 2002, 8, 4834; Rajca, A. Adv. Phys. Org. Chem. 2005, 40,
153.
[17] Rajca, A. Chem. Rev. 1994, 94, 871.
[18] Shavitt, I. Tetrahedron 1985, 41, 1531.
[19] Longuet-Higgins, H. C. J. Chem. Phys. 1950, 18, 265.
[20] Ovchinnikov, A. A. Theoret. Chim. Acta 1978, 47, 297.
[21] Borden, W. T. Ed., Diradicals, Wiley, NewYork, 1982; Borden, W. T.; Davidson, E. R.
J. Am. Chem. Soc. 1977, 99, 4587.
[22] Dougherty, D. A. Acc. Chem. Res. 1991, 24, 88; Dougherty, D. A. J. Am. Chem. Soc.
1996, 118, 1452.
[23] Klein, D. J.; March, N. H. Int. J. Quantum Chem. 2001, 85, 327.
[24] Klein, D. J.; Alexander, S. A. Stud. Phys. Theo. Chem. 1987, 51, 404; Alexander, S. A.,
Klein, D. J. J. Am. Chem. Soc. 1988, 110, 3401.
[25] Pauling, L.; Wheland, G. W. J. Chem. Phys. 1933, 1, 362.
[26] Trindle, C. O., Datta S. N. Int. J. Quantum Chem. 1996, 57, 781; Trindle, C. O.;
Datta, S. N.; Mallik, B. J. Amer. Chem. Soc. 1997, 119, 2187.
November 9, 2013 9:31 9in x 6in Theoretical and Computational Aspects. . . b1639-ch02

Organic Molecules, Radicals, and Spin States 39

[27] Datta, S. N.; Jha, P. P.; Ali, Md. E. J. Phys. Chem. A 2004, 108, 4087; Ali, Md. E.;
Datta, S. N. J. Phys. Chem. A 2006, 110, 2776, 13232; Latif, I. A.; Hansda, S.; Datta,
S. N. J. Phys. Chem. A 2012, 116, 8599.
[28] Dvolaitzky, M.; Chiarelli, R.; Rassat, A. Angew. Chem. Int. Ed. Engl. 1992, 31, 180;
Kanno, F.; Inoue, K.; Koga, N.; Iwamura, H. J. Am. Chem. Soc. 1993, 115, 847.
Theoretical and Computational Aspects of Magnetic Organic Molecules Downloaded from www.worldscientific.com

[29] Ishida, T.; Iwamura, H. J. Am. Chem. Soc. 1991, 113, 4238.
[30] Bleaney, B.; Bowers, K. D. Proc. R. Soc. London. Ser. A 1952, 214, 451.
[31] Sessoli, R.; Gatteschi, D.; Caneschi, A.; Novak, M. A. Nature 1993, 365, 141.
[32] Barra, A.-L.; Debrunner, P.; Gatteschi, D.; Schulz, C. E.; Sessoli, R. Europhys. Lett.
1996, 35, 133.
[33] Kinoshita, M.; Turek, P.; Tamura, M.; Nozawa, K.; Shimoi, D.; Nakazawa, Y.;
Ishikawa, M.; Takahashi, M.; Awaga, K.; Inabe, T.; Maruyama, Y. Chem. Lett. 1991,
by UNIVERSITY OF QUEENSLAND on 10/14/14. For personal use only.

1225.
[34] Tamura, M.; Nakazawa, Y.; Shimoi, D.; Nozawa, K.; Hosokoshi, Y.; Ishikawa, M.;
Takahashi, M.; Kinoshita, M. Chem. Phys. Lett. 1991, 186, 401.
[35] Takahashi, M.; Turek, P.; Nakazawa, Y.; Tamura, M.; Nozawa, K.; Shimoi, D.;
Ishikawa, M.; Kinoshita, M. Phys. Rev. Lett. 1991, 67, 746.
[36] Mukai, K. Bull. Chem. Soc. Japan 1969, 42, 40.
[37] Sugano, T. Polyhedron 2001, 20, 1285.
[38] Awaga, K.; Sugano, T.; Kinoshita, M. Chem. Phys. Lett. 1986, 128, 587.
[39] Awaga, K.; Sugano, T.; Kinoshita, M. J. Chem. Phys. 1986, 85, 1211.
[40] Lahti, P. M. Ed., Magnetic Properties of Organic Materials, Dekker, New York, 1999.
[41] Brown, P. J.; Capiomont, A.; Gillon, B.; Schweizer, J. J. Magn. Magn. Mater. 1979,
14, 289.
[42] Saint-Paul, M.; Veyret, C. Phys. Lett. A 1973, 45, 362.
[43] Benoit, A.; Flouquet, J.; Gillon, B.; Schweizer, J. J. Magn. Magn. Mater 1983, 31–34,
1155.
[44] Chouteau, G., Veyret-Jeandey, C. J. Physique 1981, 42, 1441.
[45] Allemand, P.-M.; Srdanov, G.; Wudl, F. J. Am. Chem. Soc. 1990, 112, 9391.
[46] Neugebauer, F. A.; Fischer, H.; Krieger, C. J. Chem. Soc. Perkin Trans. 1993, 2, 535.
[47] Mukai, K.; Konishi, K.; Nedachi, K.; Takeda, K. J. Phys. Chem. 1996, 100, 9658.
[48] Itoh, K.; Kinoshita, M. Eds, Molecular Magnetism, New Magnetic Materials, Gordon
and Breach, Amsterdam, 2000, p. 216.
[49] Mukai, K.; Suzuki, K.; Ohara, K.; Jamali, J. B.; Achiwa, N. J. Phys. Soc. Japan 1999,
68, 3078.
[50] Palacio, F.; Antorrena, G.; Castro, M.; Burriel, R.; Rawson, J.; Smith, J. N. B.;
Bricklebank, N.; Novoa, J.; Ritter, C. Phys. Rev. Lett. 1997, 79, 2336.
[51] Mito, M.; Kawae, T.; Takeda, K.; Takagi, S.; Matsushita, Y.; Deguchi, H.;
Rawson, J. M.; Palacio, F. Polyhedron 2001, 20, 1509.
[52] Segura, J. L.; Martin, N. Angew. Chem. Int. Ed. 2001, 40, 1372; Chahma, M.;
Wang, X. S.; van der Est,A.; Pilkington, M. J. Org. Chem. 2006, 71, 2750; Chahma, M.;
Macnamara, K.; van der Est, A.; Alberola, A.; Polo,V.; Pilkington, M. New J. Chem.
2007, 31, 1973; Polo, V.; Alberola, A.; Andres, J.; Anthony, J.; Pilkington, M. Phys.
Chem. Chem. Phys. 2008, 10, 857.
[53] Makarova, T. L.; Sundqvist, B.; Hohne, R.; Esquinazi, P.; Kopelevich, Y.; Scharff, P.;
Davydov, V. A.; Kashevarova, L. S.; Rakhmanina, A. V. Nature 2001, 413, 716.
November 9, 2013 9:31 9in x 6in Theoretical and Computational Aspects. . . b1639-ch02

40 Theoretical and Computational Aspects of Magnetic Organic Molecules

[54] Wood, R. A.; Lewis, M. H.; Lees, M. R.; Bennington, S. M.; Cain, M. G.; Kitamura, N.
J. Phys. Condens. Matter 2002, 14, L385.
[55] Narozhnyi, V. N.; Müller, K.-H.; Eckert, D.; Teresiak, A.; Dunsch, L.; Davydov, V. A.;
Kashevarova, L. S.; Rakhmanina, A. V. Physica B 2003, 329–333, 1217.
[56] McConnell, H. M. J. Chem. Phys. 1963, 39, 1910.
Theoretical and Computational Aspects of Magnetic Organic Molecules Downloaded from www.worldscientific.com

[57] Datta, S. N.; Misra, A.; Vinodhkumar, R. Int. J. Quantum Chem. 2000, 79, 308.
[58] Awaga, K.; Sugano, T.; Kinoshita, M. Chem. Phys. Letters 1987, 141, 540.
[59] McConnell, H. M. Proc. R. A. Welch Found. Chem. Res. 1967, 11, 144.
[60] Kinoshita, M. Phil. Trans. R. Soc. (London) A 1999, 357, 2855.
[61] Tanigaki. K.; Prassides, K. J. Mater. Chem. 1995, 5, 1515; Iwamura, H.; Inoue, K.;
Hayamizu, T. Pure Appl. Chem. 1996, 68, 243; Enoki, T.; Yamaura, J.-I.; Miyazaki, A.
Bull. Chem. Soc. Jpn. 1997, 70, 2005; Ouahab, L.; Enoki, T. Eur. J. Inorg. Chem. 2004,
by UNIVERSITY OF QUEENSLAND on 10/14/14. For personal use only.

933.
[62] Decurtins, S.; Gütlich, P.; Spiering, H.; Hauser, A. Inorg. Chem. 1985, 24, 2174.
[63] Sato, O.; Iyoda, T.; Fujishima, A.; Hashimoto, K. Science 1996, 272, 704.
[64] Enoki, T.; Yamaura, J.-I.; Miyazaki, A. Bull. Chem. Soc. Jpn 1997, 70, 2005.
[65] Kobayashi, H.; Tomita, H.; Naito, T.; Kobayashi, A.; Sakai, F.; Watanabe, T.;
Cassoux, P. J. Am. Chem. Soc. 1996, 118, 368.
[66] Coronado, E.; Galan-Mascaros, J. R.; Gimenez-Saiz, C.; Gomez-Garcia, C. J. Pro-
ceedings of the NATO Advanced Research Workshop on Magnetism: A Supramolecular
Function, Academic Press, New York, 1996.
[67] Miller, J. S.; Drillon, M. Ed., Magnetism: Molecules to Materials Vol. 1–4, Wiley–
VCH, Weinheim, 2001.
[68] de Jongh, L. J.; Miedema, A. R. Adv. Phys. 1974, 23, 1.
[69] de Jongh, L. J. Ed., Magnetic Properties of Layered Transition Metal Compounds,
Kluwer, Dordrecht, 1990.
[70] Ratera, I.; Veciana, J. Chem. Soc. Rev. 2012, 41, 303.
[71] Iwamura, H. Adv. Phys. Org. Chem. 1990, 26, 179.
[72] Buchachenko, A. L. Russ. Chem. Rev. 1990, 59, 307.
[73] Iwamura, H.; Koga, N. Acc. Chem. Res. 1993, 26, 346.
[74] McGlynn, S. P.; Azumi, T.; Kinoshita, M. Molecular Spectroscopy of the Triplet State,
Prentice-Hall, New York, 1968.
[75] Whittle, E.; Dows, D. A.; Pimentel, G. C. J. Chem. Phys. 1954, 22, 1943.
[76] Becker, E. D.; Pimentel, G. C. J. Chem. Phys. 1956, 25, 224.
[77] Norman, I.; Porter, G. Nature (London) 1954, 174, 58.
[78] Bally, T. In Reactive Intermediate Chemistry, Moss, R. A.; Platz, M. S.; Jones Jr, M.
Eds, Wiley, New York, 2004, 17, pp. 795–845.
[79] Porter, G.; Ward, B. J. The High Resolution Absorption Spectroscopy of Aromatic Free
Radicals. Chim. Phys. 1964, 1517.
[80] Iwamoto, E.; Hirai, K.; Tomioka, H. J. Am. Chem. Soc. 2003, 125, 14664.
[81] Zeng, Z.; Sung, Y. M.; Bao, N.; Tan, D.; Lee, R.; Zafra, J. L.; Lee, B. S.; Ishida, M.;
Ding, J.; López Navarrete, J. T.; Li, Y.; Zeng, W. J. Am. Chem. Soc. 2012, 134, 14513.
[82] Carlin, R. L. Magnetochemistry, Springer, Berlin, 1986.
[83] Koivisto, B. D.; Hicks, R. G. Chem. Soc. Rev. 2005, 249, 2612.
[84] Kleiner, R.; Koelle, D.; Ludwig, F.; Clarke, J. Proceedings of the IEEE 2004, 92,
1534.
November 9, 2013 9:31 9in x 6in Theoretical and Computational Aspects. . . b1639-ch02

Organic Molecules, Radicals, and Spin States 41

[85] Iwamura, H.; Koga, N.; Matsuda, K. Pure Appl. Chem. 1998, 70, 1953.
[86] Wenthold, P. G.; Lineberger, W. C. Acc. Chem. Res. 1999, 32, 597.
[87] Leopold, D. G.; Murray, K. K.; Stevens Miller, A. E.; Lineberger, W. C. J Chem. Phys.
1985, 83, 4849.
[88] Lineberger, W. C.; Borden, W. T. Phys. Chem. Chem. Phys. 2011, 13, 11792.
Theoretical and Computational Aspects of Magnetic Organic Molecules Downloaded from www.worldscientific.com

[89] Wenthold, P. G.; Kim, J. B.; Lineberger, W. C. J. Am. Chem. Soc. 1997, 119, 1354.
by UNIVERSITY OF QUEENSLAND on 10/14/14. For personal use only.

You might also like