Sunteți pe pagina 1din 11

ESSAY

DOI:10.1002/ejic.201300092

CLUSTER
ISSUE

Spin Crossover Quo Vadis?


Philipp Gtlich*[a]
Dedicated to Hartmut Spiering on the occasion of his 70th birthday
Keywords: Transition metal ions / Spin state switching / Spin transitions / LIESST / Photophysical phenomena
This retrospective essay is an attempt to span the bridge over
eight decades of research on spin crossover (SCO), one of the
most fascinating dynamic electron structure phenomena of
inorganic coordination chemistry. The occurrence of SCO
compounds of 3d transition metal ions and their characterization regarding magnetic, optical, vibrational, structural, and
thermodynamic properties are briefly addressed. Selected
case studies of chemical influences and physical effects
(pressure, magnetic field, light) on SCO behavior are discussed. Particular attention is paid to the importance of

Mssbauer spectroscopy in SCO studies. Light-induced excited spin state trapping (LIESST) and related photophysical
phenomena involving SCO compounds are briefly highlighted. The remarkable instrumental improvements over the
years and the amazing widespread developments in synthesizing new SCO compounds up to multifunctional materials
currently in the foreground of SCO research are brought to
light in a state-of-the-art summary. New directions, primarily
towards applications, are mentioned.

Occurrence of Spin Crossover

which transition metal ions are candidates capable of undergoing thermally induced SCO. There are practical rules
inferred from ligand field theory that allow predictions to
be made regarding the possible occurrence of SCO.
(1) SCO is expected to occur in octahedral complexes of
first-row transition metal ions with 47 3d valence electrons. There are practically no SCO examples known with
4d and 5d transition elements. The reason is that, in complexes of metal ions of the same group and same oxidation
state and with identical ligand sphere, the ligand field
strength increases by roughly 50 % on going from 3d to 4d
and from 4d to 5d elements, whereas the spin-pairing energy does not change much in this order. Thus, octahedral
4d and 5d complexes show a strong tendency to adopt LS
behavior.
(2) Increasing the oxidation number of an octahedral complex ion, while retaining all other features (same nd series,
same ligands), increases the ligand field parameter Dq by
about 4080 %. For instance, the ligand field strength of
Fe2+ complexes increases by about 40 % upon oxidation to
the corresponding Fe3+ complexes, while the spin-pairing
energy does not increase that much. Thus, it is not expected
that a SCO complex of iron(II) will still exhibit SCO behavior after oxidation to the corresponding iron(III) complex;
it will rather change to LS behavior.
(3) Tetrahedral SCO complexes of 3d elements are not
known, because the ligand field strength is only about half
that of octahedral complexes,[5] thus favoring HS behavior.
(4) Among the octahedral 3d transition metal complexes
exhibiting SCO, those with 5, 6, and 7 3d electrons show

Thermally induced spin crossover (SCO), also known as


spin transition (ST), is one of the most fascinating dynamic
electronic structure phenomena occurring in inorganic coordination chemistry. Its roots date back by more than eight
decades to the early 1930s, when Cambi and co-workers
reported on the observation of unusual magnetic properties
of dithiocarbamato complexes of iron(III).[13] They interpreted, erroneously, the observation of temperaturedependent moments as arising from equilibria between two
magnetic isomers. Nonetheless, their pioneering work has
opened the door to a fascinating research area of coordination chemistry that has attracted the interest of chemists
and physicists ever since. With the development of ligand
field theory (LFT),[4,5] commonly denoted as crystal field
theory in solid-state physics, a theoretical model has been
introduced that, in conjunction with group theoretical
methods, has most successfully been applied for the chemical and physical characterization (magnetism and spectral
properties) of transition metal complexes. In particular, the
principle of the SCO phenomenon, and the thermal condition to be fulfilled in order to observe it, are now well
understood.[6,7]
After Cambis discovery of spin state interconversion in
iron(III) complexes, the immediate question arose as to
[a] Institute of Inorganic and Analytical Chemistry,
Johannes Gutenberg University Mainz,
Staudingerweg 9, 55099 Mainz, Germany
E-mail: guetlich@uni-mainz.de
Homepage: www.ak-guetlich.chemie.uni-mainz.de
Eur. J. Inorg. Chem. 2013, 581591

581

2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.eurjic.org
a comparatively strong tendency to undergo thermal spin
transition because of favorable competition between the ligand field stabilization energy and the mean spin-pairing
energy (comprising Coulomb repulsion and quantum-mechanical exchange energy).[5,8] In fact, most SCO complexes
known are complexes of iron(III), iron(II), and cobalt(II).
Only few SCO complexes of manganese(II), manganese(III), chromium(II), and cobalt(III) have been reported.[9]

SCO Complexes A Brief Survey


For nearly three decades after Cambis discovery, synthetic work and physical characterization was reported solely on iron(III) SCO systems. Only in the 1950s did the
first examples of iron(II) and cobalt(II) exhibiting thermally induced SCO become known. Interest in this fascinating electronic switching phenomenon has rapidly increased,
and up to now several hundred SCO compounds have been
synthesized and characterized, by far the largest portion belonging to SCO systems of iron(III), iron(II), and cobalt(II). Within the scope of the present essay there is room
only for a limited coverage of selected examples without
going into detail. The reader is advised to consult the first
comprehensive account on SCO research published in
2004[10] and other more recently published relevant articles
cited below.

SCO Complexes of Iron(III)


Van Koningsbruggen et al. have extensively reviewed the
literature of SCO compounds of iron(III) up to 2003.[11]
Their overview of the large variety of prepared and investigated iron(III) SCO complexes is categorized with respect
to the donor atom sets:
Iron(III) SCO complexes with chalcogen donor atoms:
tris(N,N-disubstituted dithiocarbamato) compounds (S6);
tris(N,N-disubstituted XY-carbamato) compounds (XY =
SO, SSe, SeSe);
tris(substituted X-xanthato) compounds (X = O, S);
tris(monothio--diketonato) compounds;
bis(X-semicarbazone) compounds (X = S, Se).
Iron(III) SCO complexes with multidentate ligands of
the Schiff base type:
complexes with tridentate N2O-donating ligands;
complexes with tetradentate N4 ligands;
five-coordinate complexes with tetradentate N2O2-donating ligands;
six-coordinate complexes with tetradentate N2O2-donating
ligands;
complexes with pentadentate N3O2-donating ligands;
complexes with hexadentate N4O2-donating ligands.
Attention should also be paid to the selection of more
recent work on SCO iron(III) compounds, as examples of
the growing widespread research interest in this area, in
refs.[1232]
Eur. J. Inorg. Chem. 2013, 581591

ESSAY

Six-coordinate SCO complexes of iron(III) generally exhibit S = 1/2 S = 5/2 spin transitions, whereas for fivecoordinate systems an intermediate S = 3/2 S = 5/2 spin
transition is observed for a fairly large variety of metal-ionto-donor-atom sets: FeOS4, FeNS4, FeN3O2, FeN5, Fe4NX
(X = Cl, Br).[11] The spin transition behavior, expressed in
terms of the molar fraction of HS molecules as a function
of temperature, HS(T), is generally rather gradual, as compared to SCO systems of iron(II), and hysteretic behavior
has hardly been observed. Both features are due to the lack
of significant cooperative interactions. This is different in
the case of SCO complexes of iron(II), as will be pointed
out in the next section.

SCO Complexes of Iron(II)


It was only in the early 1960s, about three decades after
Cambis discovery of the first spin state interconversion in
iron(III) complexes, that the SCO phenomenon was reported to occur also in iron(II) compounds.[3335] The first
SCO compounds of iron(II) described in the literature are
[Fe(phen)2X2] (phen = 1,10-phenanthroline; X = NCS,
NCSe), frequently called the classical iron(II) SCO complexes. Extensive research activities began thereafter in this
area, particularly favored by the discovery of the Mssbauer effect (recoilless nuclear resonance absorption) by Rudolf L. Mssbauer[36,37] at nearly the same time. This nuclear resonance effect formed the basis for the development
of a new spectroscopic technique, Mssbauer spectroscopy,
for the detection of hyperfine interactions,[38] which turned
out to be extremely powerful for the characterization of
iron-containing substances.[39]
SCO occurs mainly for six-coordinate iron(II) compounds involving the change of electron configuration
t2g6eg0 (1A1g, LS) t2g4eg2 (5T2g, HS). Many review articles
describing the preparation, structure, chemical, and physical properties of SCO systems of iron(II) have appeared in
the literature.[7,10,4051] Most iron(II) SCO systems possess
an [FeN6] coordination center, but examples with other donor-atom sets have also been reported, for example,
N4O2[5254] and P4Cl2.[55] Various classes of iron(II) SCO
compounds have been reviewed in ref.:[10]
tris(diimine) and bis(terimine)systems;[56]
pyrazolylborate and pyrazolylmethane systems;[57]
complexes of 1,2,4-triazole, isoxazole, and tetrazole ligands;[58]
complexes with multidentate ligands (tetradentate N4,
pentadentate N5, hexadentate N4O2, hexadentate N6, heptadentate N7, octadentate N8);[59]
bipyrimidine-bridged dinuclear compounds;[60]
1D, 2D, and 3D polymeric networks.[61]
Iron(II) SCO compounds are the most extensively
studied of all SCO systems for several reasons: (1) In addition to generally employed methods like magnetic susceptibility measurements and optical spectroscopy for the fundamental characterization of spin transition behavior,
Mssbauer spectroscopy works very well, particularly for

582

2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.eurjic.org
iron(II) compounds; in most cases it yields well resolved
sharp spectra, whereas in the case of iron(III) SCO compounds the resonance signals are often very broad and difficult to analyze for reasons that are pointed out below. (2)
Spin transition curves HS(T), which are highly informative
for SCO behavior, can adopt a variety of forms for systems
in the solid state, gradual, abrupt, stepwise, with or without
hysteresis, even in the room temperature regime.[7] The latter feature is a prerequisite for bistability, which makes such
materials attractive for possible applications. (3) Iron(II)
SCO compounds have turned out to be very favorable for
studies of cooperative interactions. The material can be
chemically modified in various ways, which effectively influence the SCO behavior as signalized by the shape of the
spin transition curves. More about this important aspect
will be discussed below under Chemical Influences on
SCO Behavior.

SCO Complexes of Cobalt(II)


The SCO phenomenon is much less frequently observed
for cobalt(II) than for iron(II) complexes. The first characterization of a spin transition in a six-coordinate cobalt(II)
compound by Stoufer et al.[62] may be regarded as the beginning of SCO research of cobalt(II) complexes. The majority of SCO systems of cobalt(II) that are known are sixcoordinate compounds involving thermally induced spin
transition with the change of electron configuration t2g6eg1
(2Eg, LS) t2g5eg2 (4T1g, HS). Such changes of spin state
are also observed for four-coordinate and some five-coordinate complexes. In these cases, however, the phenomenon is
more likely assigned to configurational equilibria.[63] Goodwin has reviewed the literature on SCO complexes of cobalt(II)[64] comprising:
bis(tridentate) systems;
tris(diimine) systems;
tetradentate Schiff base systems;
five-coordinate [Co(tridentate)X2] systems;
configurational equilibria;
trinuclear systems.
Attention is also drawn to reports that have appeared
more recently.[18,23,6568] The review of Krivokapic et al.[66]
presents in-depth case studies of tris(bipyridine) and bis(terpyridine) complexes of cobalt(II), discussing results from
magnetic measurements, crystal structure determination,
optical and EPR spectroscopy on the grounds of ligand
field theory, symmetry lowering, and JahnTeller effect occurring in the LS state 2Eg.
Spin transitions of iron(II) and cobalt(II) compounds
show similar features in some cases but also distinct differences in others, most likely due to several factors: (1) the
smaller change of spin multiplicity {S = 1 for cobalt(II),
whereas S = 2 for iron(II)} and the concomitant smaller
change of the distance between the metal ion and the donor
atom for CoII than that for FeII SCO systems; (2) influences
of the JahnTeller effect that is operative in the LS state
of six-coordinate cobalt(II), leading to additional molecular
Eur. J. Inorg. Chem. 2013, 581591

ESSAY

distortion, which is not present in corresponding iron(II)


complexes; (3) the larger mean spin-pairing energy for CoII
relative to that for FeII in comparable complexes, which requires higher ligand field strength to effect spin pairing and
thus spin transition to the LS state in CoII SCO systems. A
remarkable difference between the SCO behavior of cobalt(II) and iron(II) complexes is that hysteretic spin transition behavior is hardly observed in cobalt(II) complexes.

SCO in Other Transition Metal Ions


Occurrence of SCO in other 3d transition metal complexes is extremely scarce. Some examples were observed
for 3d4 compounds of chromium(II) and manganese(III)
with the t2g4eg0 (3T1g, LS) t2g3eg1 (5Eg, HS) transition
(under Oh symmetry), for 3d5 compounds of manganese(II)
with the t2g5eg0 (2T2g, LS) t2g3eg2 (6A1g, HS) transition,
and for 3d6 cobalt(III) with the t2g6eg0 (1A1g, LS) t2g4eg2
(5T2g, HS) transition; these have been reviewed by Garcia
and Gtlich.[9] Later on, a few more reports that describe
the occurrence of SCO triggered by geometric tuning in
manganese(III)[69] and the synthesis, structural, and magnetic characterization of a new manganese(III) compound
with a MnN4O2 chromophore[70] appeared.
An interesting case is the so far only observation of thermally induced SCO in a cobalt(III) coordination compound
of formula [CoL2]PF6 by W. Klui,[7173] where the central
CoIII ion is octahedrally coordinated through oxygen atoms
of two tridentate tripodal ligands with L = {(C5H5)Co[P(O)(OC2H5)2]3}; the CoIII atoms in the tripodal ligands
are diamagnetic, but the central CoIII exhibits a gradual
temperature-dependent spin transition. This situation is entirely unexpected, because it is well known that practically
all cobalt(III) coordination compounds are LS, except for
complexes [CoF6] and [CoF3(H2O)3], which are HS, and
complex [Co(H2O)5(OH)]2+, which is known to be LS. Apparently, as inferred from the spectroscopic series of ligands,
there must be a marked difference in ligand field strength
between the strongly electronegative F ligands and the less
electronegative oxygen atoms of the directly neighboring
H2O and OH ligands. Obviously, in the present case the
ligand field strength acting at the central cobalt(III) ion is
fine-tuned by the phosphorus atoms in the next ligand
sphere to the critical value where thermal SCO sets in. The
spin transition curve was followed in solution with the
NMR spectroscopic method of Evans referring to 31PF6 as
standard.[74]
Spin state equilibria in solution have also been described
for nickel(II) (3d8) complexes.[9] However, in all these cases,
changes of spin states arise from configurational changes
such as tetrahedral HS square planar LS. It is inferred
from ligand field theory that thermally induced spin transitions originating from the competition between ligand field
strength and spin-pairing energy should be possible, but, to
our knowledge, no such case has been communicated so far
for nickel(II). In this context, the recent work of Tuczek,
Herges et al. on coordination-induced SCO, abbreviated as

583

2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.eurjic.org
CISCO, through axial bonding of substituted pyridines to
nickel(II)porphyrin complexes,[75] should be pointed out.
Along this line, the same research group has proposed a
new principle of magnetic switching of molecules based on
light-driven coordination-induced spin state switching (LDCISSS) between square-planar LS (S = 0) and octahedral
HS (S = 1) nickel(II)porphyrin complexes in solution in
the room temperature region.[76,77]

Physical Characterization of SCO Complexes


Thermally induced SCO in coordination compounds is
accompanied by changes of various molecular properties,
magnetic, optical, vibrational, structural, and thermodynamic changes being the most important ones. Measuring the magnetic susceptibility as a function of temperature,
(T), has been a standard method from the very beginning
of SCO research, because the change in the number of unpaired electrons between the HS and LS states is readily
reflected in a drastic change of (T). The spin transition
curve expressed in terms of the molar fraction of HS molecules, HS(T), can be derived from the equation (T) =
HS(T) HS + [1 HS(T)] LS. Spin transition curves of SCO
in solution are always gradual and can be described by a
Boltzmann distribution over all vibronic energy levels involved. In the solid state, however, spin transition curves
can adopt a variety of shapes: gradual, abrupt, with one or
more steps, with a hysteresis of different widths and positions on the temperature scale, and they can be incomplete
at higher and at lower temperatures.[7] This diversity arises
from cooperative interactions between the spin state changing complex molecules, which do not play an important role
in the solution state. It is the existence of cooperativity in
the solid state, which, on the one hand, makes the mechanism of SCO processes in solids so difficult to understand
and to develop appropriate theoretical models for, but, on
the other hand, is very stimulating and attractive to chemists and physicists in view of promising technical applications.
SCO is always accompanied by a color change. Therefore, optical spectroscopy in the UV/Vis range on single
crystals is often employed in SCO studies; it is also suited
to track the spin transition curves HS(T), provided the ligand field (dd) absorption bands of interest are well resolved and not hidden under the much stronger chargetransfer bands. Optical spectroscopy is, of course, most important for studies of photophysical processes such as
LIESST (see below).
In the course of spin transition from the HS to the LS
state in octahedral SCO complexes, there is charge depletion in the antibonding eg orbitals and an increase of
charge population in the slightly bonding t2g orbitals. This,
of course, strengthens the metal-ion-to-donor-atom bonds
considerably. For instance, the stretching frequency observed in the far infrared region changes from typically
300 cm1 in the HS state to about 450 cm1 in the LS state.
Both FIR and Raman spectroscopy are employed in SCO
Eur. J. Inorg. Chem. 2013, 581591

ESSAY

studies to follow the temperature-dependent intensities,


decreasing for typical HS and increasing for LS vibrational
bands, upon lowering the temperature. Analysis of vibrational spectra and band assignment are often cumbersome because of the complexity of vibrational spectra with
a great variety of many bands. And yet, it is possible to
trace the spin conversion curves, HS(T), in favorable cases
by determining the area fractions of well resolved and isolated characteristic HS or LS bands. A new method of
studying vibrational properties of SCO complexes has been
developed on the basis of nuclear resonant scattering of
synchrotron radiation.[78] The method makes use of the
Mssbauer effect (in the time domain).[38] The sample under study containing a Mssbauer-active isotope, for example, 57Fe, is irradiated with sharply tuned synchrotron radiation of energy around 14.4 keV, the energy of the first excited state of 57Fe. The precise nuclear excitation energy is
14412.5 eV; this is absorbed elastically out of the radiation
beam with a certain energy width and appears as an elastic
or zero-phonon peak in the absorption spectrum. Simultaneously, inelastic absorption of synchrotron radiation of the
same beam occurs and excites molecular vibrations in the
near neighborhood of the 57Fe central ion. The resulting
inelastic phonon peaks appear as sidebands next to the elastic nuclear absorption peak in the FIR energy range, characteristic of metalligand bond vibrational frequencies of
HS and LS states. This technique is called nuclear inelastic
scattering of synchrotron radiation (NIS), also known as
phonon-assisted Mssbauer effect or nuclear resonance vibrational spectroscopy. The great advantage of NIS is that
it is site-selective, that is, the phonon peaks displayed in the
vibrational spectrum originate only from atomic vibrations
in the near neighborhood of Mssbauer-active nuclides,
which makes the vibrational spectrum much less complicated than FIR and Raman spectra. In the case of 57Fe
experiments, vibrational frequencies ranging down to about
20 cm1 can be covered with NIS, with an energy resolution
of about 4 cm1. Applications in SCO research are described in the literature.[7981]
For the characterization of iron-containing SCO compounds, 57Fe Mssbauer spectroscopy[38] has become an
eminently important method. There is hardly any report
dealing with studies of SCO in iron compounds where this
technique has not been used. As a microscopic tool, it readily provides information on oxidation and spin states, magnetic behavior, molecular symmetry and distortions, and
lattice dynamics to name a few important properties,
through hyperfine interaction measurements. For the characterization of SCO compounds, the isomer shift and electric quadrupole splitting are the most important parameters;
they differ significantly between HS and LS states, both in
iron(II) and iron(III) compounds. From the area fractions
of the resonance signals, which are approximately proportional to the iron concentrations of the coexisting spin
states (if minor differences in the so-called LambMssbauer factors are neglected), the spin transition curves
HS(T) can easily be plotted. With Mssbauer spectroscopy,
it is also possible to determine the rate of spin state switch-

584

2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.eurjic.org
ing, kHL, from line-shape analysis of relaxation-broadened
resonance lines with especially developed relaxation models.
The time window of Mssbauer measurements is given by
the lifetime of the nuclear excited state involved, for example, for 57Fe spectroscopy n 100 ns. In the case of interconverting HS and LS spin states, sharp and well resolved resonance lines are observed if the interconversion
rate, kHL, is sufficiently slow compared to the reciprocal of
the nuclear excited state lifetime (kHL 1/n). This is generally the case with SCO compounds of iron(II). However,
in the case of SCO compounds of iron(III), the interconversion rate, kHL, is faster than 1/n, and only time-averaged
species are observed. In cases where the interconversion rate
becomes comparable to 1/n, temperature-broadening of the
resonance lines is observed, which can be used in conjunction with special relaxation models to determine the interconversion rate, kHL, and thus the lifetimes, 1/kHL, of the
switching spin states involved.
As already mentioned, SCO occurring in solid materials
is accompanied by more or less significant structural
changes because of the drastic electron rearrangement between the different spin states involved. It is therefore indispensable to include structural investigations in studies of
SCO phenomena, mostly by X-ray diffraction on single
crystals or polycrystalline materials (PXRD).[82,83] SCO
compounds with continuous spin transition generally show
only changes of metaldonor atom bond lengths and
angles, but no change of space group (displacive transitions). Changes in the lengths of the bonds between the
metal and the donor atom are about 10 % (rHL = rHS
rLS 220200 pm) in iron(II) SCO systems with = 2 transitions, about 1013 pm in iron(III) SCO systems with =
2 transitions, and less than or equal to 10 pm in cobalt(II)
systems with = 1 transitions. Discontinuous spin transitions with hysteretic behavior in the transition curve HS(T)
show in most cases a change of space group, signalizing the
occurrence of a first-order phase transition with significant
lattice reorganization (reconstructive transitions). In order
to identify the type of spin transition, it is essential to carry
out structural investigations above and below the spin transition temperature T1/2 (temperature of half-way spin transition, where 50 % of the SCO-active molecules have
changed their spin state; this formulation also refers to incomplete spin transitions with residual paramagnetism at
lower temperatures and/or residual LS behavior in the room
temperature region or above). Structural changes accompanying continuous and discontinuous spin transitions
have been extensively reviewed by Knig.[84] Detailed
knowledge about structural changes going along with spin
transition in solids is of utmost importance for in-depth
studies of cooperative interactions and SCO mechanisms.
These have been among the major objectives of SCO research ever since it has been realized that cooperative interactions are present[85] and do play a decisive role in spin
transition processes in solid materials, particularly in view
of possible applications. It is worth referring in this context
to a recently published article by Halcrow[86] on structure
function relationships in molecular SCO complexes. InstruEur. J. Inorg. Chem. 2013, 581591

ESSAY

mentation for structural studies has continuously been improved, both in university laboratories as well as at research
centers providing neutron diffraction and synchrotron radiation facilities. Sophisticated equipment is nowadays available for structural studies at cryogenic temperatures and on
metastable spin states,[83] even on very small sample sizes.
Calorimetric measurements (DSC, precise heat capacity
measurements) yield information on thermodynamic quantities such as enthalpy and entropy changes and the order
of phase transitions. Knowledge on these features can be
very helpful in clarifying mechanistic aspects of SCO[7,8789]
and are therefore often included in SCO studies. Thermally
induced SCO is an entropy-driven process, because the degree of freedom is considerably higher in the HS than in
the LS state. The total entropy gain connected with a spin
state conversion from LS to HS is composed of two contributions. The smaller one, approximately 25 % of the total
entropy gain of typically 5080 J mol1 K1, arises from the
change in spin multiplicity, Smag = R[ln(2S + 1)HS ln(2S
+ 1)LS], and amounts to about 13 cal mol1 K1 in the case
of SCO complexes of iron(II). The major part originates
from intramolecular, and to a lesser extent from intermolecular, vibrations.[90,91]
These physical methods are employed in most SCO studies on solid materials as standard techniques. There are a
few more techniques, which are less frequently used, such
as magnetic resonance spectroscopy (EPR, NMR),[74,9294]
X-ray absorption spectroscopy (EXAFS, XANES),[9597]
positron annihilation,[98] muon spin rotation.[99,100] These
techniques are applied under special experimental conditions or if certain information is not accessible otherwise.

Influences on SCO Behavior


Chemical Influences on SCO Behavior
Already in the early stage of SCO research, it was recognized that SCO behavior can vary considerably with the
nature of the coordinated ligands, as reflected by the shape
of the spin transition curve, HS(T). Strategies of preparative chemistry, often involving advanced techniques of synthetic organic chemistry, have been developed in order to
alter the ligand sphere in a more or less controlled manner,
aiming at the desired SCO behavior, generally in two ways:
(1) ligand exchange, (2) intraligand substitution. An example of the former is replacing one phen ligand in the LS
complex [Fe(phen)3]2+ by two NCS ions to yield the SCO
complex [Fe(phen)2(NCS)2]. An example of the latter type
is the preparation of the SCO complex [Fe(2-CH3
phen)3]2+ by substitution of H by CH3 in the 2-position of
the LS complex [Fe(phen)3]2+. Many case studies in this
regard have been reported.[7,40,41] It is obvious that these
changes of the chemical composition influence the ligand
field strength acting at the central metal ion in several ways:
electronic effects directly influence the basicity of the coordinating donor atom and thus the -interaction between
metal and donor atom; steric hindrance by rotating substituents close to the donor atom influences the distance, r,

585

2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.eurjic.org
between the metal and the donor atom and thereby the ligand field strength according to Dq 1/r5;[5] hydrogen
bonding interconnecting spin state changing complex molecules and stacking of ligands involving aromatic components are known to exercise considerable influence on the
SCO behavior; finally 1D, 2D, and 3D polymerization tend
to increase the cooperative interactions and are often the
source of abrupt transitions with hysteresis. It is almost superfluous to mention that these influences operate in a concerted manner and that it is difficult or even impossible to
make reliable predictions of the effectiveness of the various
contributions. It is, therefore, not appropriate to speak of
fine-tuning the ligand field when preparing solid SCO
materials.
Influences of non-coordinated anions and solvent molecules have also been well established and documented.[7,101]
Isotope-effect studies with H/D- and 14N/15N-labeled SCO
compounds have been performed to explore hydrogen
bonding networks and stepwise spin transitions.[7,102]
Studies of the SCO behavior in isomorphous mixed crystals, prepared by gradually substituting the SCO metal ion
by a different transition metal ion, were first carried out by
Sorai et al. on the series [FexZn1x(pic)3]Cl2EtOH (pic =
picolylamine) by using 57Fe Mssbauer spectroscopy.[85] It
was found that the spin transition curve HS(T), being
rather steep for the pure iron(II) compound (x = 1), became
more and more gradual with decreasing iron concentration.
In highly diluted mixed crystals with only about 1 % or less
iron(II) (using samples with enriched 57Fe), the HS(T)
curve could be described with the Boltzmann distribution
function, similar to the general SCO behavior in liquid
solution, where cooperative interactions are practically nonexistent. These findings have been confirmed in similar
metal-dilution studies and clearly indicate that cooperative
interactions are of utmost importance in solid SCO materials. A theoretical approach based on the results of these
experiments was developed by H. Spiering of the Mainz
group, featuring lattice expansion, image pressure, and elastic interactions accompanying SCO processes in solids.[7,103105] This so-called Mainz Model has successfully
been employed to describe experimental HS(T) curves, even
the occurrence of hysteretic behavior.

Physical Influences on SCO Behavior


The influence of pressure on SCO behavior has always
attracted the interest of researchers.[106110] In view of the
marked shortening of the distance between metal and donor atom resulting from the electron rearrangement upon
going from the HS to the LS state, it is expected that the
LS state of a SCO compound is generally stabilized, which
means that the HS(T) curves tend to be shifted upwards on
the temperature scale, with or without a change in the shape
of the transition curve; examples are discussed in ref.[109]
It has been realized that the application of relatively low
hydrostatic pressure is more informative and therefore more
appropriate than the application of high pressure with diaEur. J. Inorg. Chem. 2013, 581591

ESSAY

mond anvil devices, because subtle effects, for example, on


transition steps[111] or hysteresis loops,[112] may be easily
blurred under high pressure. Specially designed hydrostatic
pressure cells for magnetic susceptibility measurements with
a SQUID magnetometer and for Mssbauer effect measurements at variable temperature down to about 4 K and at
pressures up to about 14 kbar have been developed.[109]
In view of the involvement of strongly paramagnetic species in SCO phenomena, it has been speculated that application of a magnetic field should somehow influence the
SCO behavior. A theoretical estimate showed that the effect
should only be minor, for example, the transition temperature T1/2 is expected to shift by about 0.1 K in a field of
5 Tesla.[113] This has indeed been found with measurements
on [Fe(phen)2(NCS)2], which yielded 0.11 0.04 K.[113]
Later on, Bousseksou et al. carried out similar measurements employing considerably higher magnetic fields, static
and pulsed, and explored the influence of the magnetic field
in great detail.[114]

Photophysical Phenomena Involving SCO


Compounds
Relaxation kinetics and lifetime measurements on excited
ligand field states of coordination compounds in solution
have been extensively studied by using ultrasonic, laser-temperature-jump, and direct laser excitation techniques.[115] In
connection with time-differential Mssbauer emission spectroscopy (TDMES)[116] of excited ligand field states of 57Fe
resulting from the nuclear decay of 57Co, embedded in the
LS compound [Fe(phen)3](ClO4)2 and used as Mssbauer
source, it was attempted, for comparison, to measure the
lifetime of the photoexcited quintet ligand field state 5T2 of
[Fe(phen)3]2+ in a frozen solution of an ethanol/methanol
mixture (4:1) after direct laser population of the 1MLCT
state and decay by double intersystem crossing to the 5LF
state.[115e] Shortly afterwards, an entirely unexpected observation was made by the Mainz group by irradiation of a
single crystal of the SCO compound [Fe(ptz)6](BF4)2 (ptz =
1-propyltetrazole). It was observed for the first time that
the spin transition from LS(1A1) to HS(5T2) can be induced
by light; green light converts the LS state to a long-lived
metastable HS state, which may have lifetimes of the order
of hours or longer at cryogenic temperatures.[117] This phenomenon, known as light-induced excited spin state trapping
(LIESST), has opened a fascinating new direction of SCO
research.[117] Later it was found that reverse-LIESST, that
is, pumping the metastable HS state back to the LS state,
is also possible by irradiating the crystal into the HS ligand
field band in the near infrared.[118] LIESST and reverseLIESST are well understood on the basis of ligand field
theory.[7,117b,118b,119] Light-induced SCO and HSLS relaxation were examined experimentally and theoretically by
Hauser.[119] He interpreted this phenomenon with a nonadiabatic multiphonon relaxation model, where the lifetime
of the metastable LIESST state, that is, the inverse of the
low-temperature tunneling rate tHL0 = (kHL0)1, is corre-

586

2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.eurjic.org
lated with the energy difference, EHL0, between the lowest
vibronic energy levels of the HS and LS states involved.
EHL0 increases with increasing ligand field strength, which
means that, at comparable temperatures, the weaker the ligand field strength is, the longer is the lifetime of the
LIESST state (inverse energy gap law).[119] The very short
lifetimes, of the order of a few nanoseconds, of the 5LF
state found in the above-mentioned laser-induced spin state
conversion experiment with the LS complex [Fe(phen)3]2+
in a frozen solution,[115e] in comparison with the much
longer LIESST state lifetimes of the order of hours or
longer, observed in SCO compounds of iron(II), are well
rationalized with this model. However, entirely unexpectedly and contrary to the inverse energy gap law, extremely
long lifetimes have been observed in a LS complex of
iron(II); the phenomenon was named strong-field
LIESST.[120] It was first believed that LIESST phenomena
are only observable with SCO compounds of iron(II) because of their favorable so-called HuangRhys factors arising from the relatively large difference in metaldonor-atom
distances between HS and LS states.[119] In fact, LIESST
phenomena have so far mostly been observed in SCO compounds of iron(II),[121] in mononuclear, as described above,
dinuclear,[122] polymeric systems,[123] and bifunctional systems like mesogens.[124] Examples of the occurrence of
LIESST in iron(III) compounds[21] have also been reported.
In view of eventual applications it would, of course, be
highly desirable to have material with long-lived metastable
LIESST states. Ltard et al. have proposed guidelines to
the design of such materials.[121b] The competition between
relaxation and photoexcitation in solid SCO compounds
under continuous irradiation, as well as the occurrence of
photomagnetic effects related to LIESST, such as light-induced thermal hysteresis (LITH)[125] and light-induced optical hysteresis (LIOH), has been treated by Varret et al.[126]
It has been found that LIESST is not only effected by
irradiation with visible light. Renz et al. have reported on
hard-X-ray-induced excited spin state trapping (HAXIESST);[127] they also showed that continuous irradiation
of [Fe(phen)2(NCS)2] can lead to what they called hardX-ray-induced thermal hysteresis (HAXITH) in a SCO
compound.[128]
A most unusual LIESST-like phenomenon is nuclear decay-induced excited spin state trapping (NIESST).[116] This
technique is based on Mssbauer emission spectroscopy
(MES), where, in the case of 57Fe spectroscopy, the compound under study is labeled with radioactive 57Co and
used as the Mssbauer source against a reference iron compound as absorber. The energy released by the electron capture (EC) nuclear decay, 57Co(EC)57Fe, is transferred to the
nearby ligand sphere and leads, among various aftereffects, to very short-lived different excited ligand field
states, for example, spin-singlet, -triplet, and -quintet states.
These relax either promptly (spin allowed) or by intersystem crossing to the stable LS ground state or to the metastable NIESST state, following the same mechanism as in
LIESST relaxation. The branching ratio depends on the ligand field strength of the iron compound corresponding to
Eur. J. Inorg. Chem. 2013, 581591

ESSAY

the 57Co labeled source compound. For instance, using


57
Co labeled strong-field compounds like the LS compound
[Fe(phen)3](ClO4)2 as Mssbauer source leads to much
shorter lifetimes of the NIESST state than with 57Co labeled intermediate-field SCO compounds as Mssbauer
source.[116] The nuclear decay serves as an intramolecular
light source in NIESST experiments.
A fascinating LIESST-like phenomenon, namely spin
state switching of a single molecule, has recently been communicated by Gopakumar, Tuczek et al.[129] The authors
prepared a double-layered sample of [Fe(bpz)2phen] on
gold {bpz = dihydrobis(pyrazolyl)borate} and scanned it at
4 K with a scanning tunneling microscope. Electrons are
thereby injected through the STM tip and induce controlled
reversible spin state switching between the stable LS state
and the metastable HS state. The authors call this phenomenon, which is highly attractive for possible applications,
electron-induced excited spin state trapping (ELIESST).
LIESST and related phenomena discussed so far have
in common that the central metal ion of the coordination
compound under study is directly addressed by external or
internal (in the case of NIESST) stimuli. On the contrary,
Boillot and Zarembowitch have proposed and successfully
pursued an indirect strategy of spin state switching: liganddriven light-induced spin change (LD-LISC).[130] This photomagnetic effect is based on the modulation of the ligand
field strength through a photochemical reaction, for example, light-induced cis/trans-isomerization, on the ligand.
LD-LISC has been performed with visible light on iron(II)
and iron(III) compounds, either in solution or embedded in
polymeric matrices.

State of the Art Future Directions

587

The strongly limited scope of an essay like this has not


allowed to cover all highlights and landmarks of eight decades of SCO research that would have deserved to be dealt
with in this account. Sincere apologies are therefore extended to all those who have contributed praiseworthy work
on SCO studies and feel that they have not been adequately
discussed or cited in this essay.
Remarkable progress in making new SCO compounds
with desired features, in investigating their properties by
physical means and in trying to understand the mechanisms
of spin transition has only been possible by extensive collaboration of chemists, physicists, and theoreticians. Physical characterization methods have continuously been improved or new ones invented. For instance, from the weighing methods of magnetic susceptibility (Gouy balance and
Faraday balance) in the early stages of SCO studies, developments went on to vibrating sample magnetometers
(Foner), to AC and DC methods, and led, after the discovery of the Josephson effect, to the nowadays most commonly used SQUID techniques. Instrumentation for structural studies has experienced remarkable improvement, too,
particularly for single-crystal diffraction measurements at
cryogenic temperatures, even on photoinduced metastable
2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.eurjic.org
spin states. Access to synchrotron radiation facilities has
immensely stimulated structural studies; fine-tuned synchrotron radiation with high brilliance and time structure
enables diffraction, EXAFS, and XANES measurements on
tiny small crystals. The invention of nuclear resonance scattering (NRS) techniques, NIS and NFS, have opened the
door to information that was not accessible otherwise: (1)
nuclear inelastic scattering of synchrotron radiation (NIS)
for measurements of local phonon densities of vibrational
states, yielding much less complicated vibrational spectra
than IR and Raman spectra, and (2) nuclear forward scattering (NFS) for Mssbauer effect measurements in the
time domain, yielding hyperfine interaction parameters like
isomer shift and quadrupole splitting as in conventional
Mssbauer spectroscopy, but now also feasible with nonclassical Mssbauer-active nuclides. These synchrotron radiation techniques, though only recently introduced in SCO
research, surely will play an important role in future SCO
studies. Remarkable advances in laser technology, time-dependent optical and vibrational spectroscopy, have made it
possible to study ultrafast processes during spin transition,
occurring in the sub-picosecond and femtosecond regime.[131] Most spectacular is the recently reported successful single-molecule switching of spin states with custombuilt STM;[132] this represents a real breakthrough with
high potential for future applications. This list of instrumental achievements in SCO research could be extended
further. An important part of the physical characterization
concerns the electronic structure of open-shell transition
metal compounds. It is therefore indispensable to add advanced electronic structure calculations, which has been
carried out successfully using DFT methods.[133]
Amazing advances have also been accomplished in the
preparation of SCO materials, mainly owing to the availability of elaborate strategies and synthesis methods developed by organic chemists for the design and synthesis of
organic ligand molecules. SCO research commenced with
mononuclear coordination compounds in Cambis early
work; they maintained their attractiveness for fundamental
studies up to now with SCO compounds of different 3d
transition metal ions, mostly FeII, FeIII, and CoII, and ligands of different denticity. Dinuclear SCO compounds of
FeII were added in the 1980s,[60,134] followed by trinuclear,[135] tetranuclear,[136] and higher-nuclearity[137] SCO
systems of FeII. The era of polymeric SCO materials with
1D, 2D, and 3D networks of systems of the triazole, tetrazole, and pyridine type began with the pioneering work of
Haasnoot[138] and Lavrenova,[139] which has spontaneously
stimulated extensive SCO studies by many other groups because of the favorable SCO properties of polymeric materials for molecular electronics devices.[61,140] It has been recognized that the appearance of hysteresis in the spin transition curve, preferentially in the room-temperature region
with broad width and sharp descending and ascending
branches of a hysteresis loop, results from cooperative interactions and generates bistability as the most important prerequisite for technical applications. Kahn was probably the
first to point out that spin transition polymers bear the poEur. J. Inorg. Chem. 2013, 581591

ESSAY

tential for applications in memory devices.[141] Ltard and


co-workers thereafter demonstrated the feasibility of this
application with some examples.[142] Undoubtedly, future
activities in SCO research, encompassing efforts in all disciplines like preparation, characterization, and theoretical
treatment of SCO materials, will be conducted extensively
towards applications. Interestingly, in nearly all recent and
current publications dealing with SCO studies the authors
emphasize this aspect. Clearly, the current state of the art
signalizes already remarkable progress in this endeavor, as
can be seen in the large variety of special SCO systems such
as thin films,[143] gels,[144] nanoparticles,[145] cages,[146]
micro/nanoporous[147] materials, and multifunctional SCO
systems[148] that are currently under extensive exploration.
Study of the latter group of materials, combining SCO behavior with other physical phenomena such as liquid crystalline properties,[49,149] electrical conductivity,[150] fluorescence,[151] to name a few, is still in an early stage of pioneering work, but enjoys much attractiveness in view of possible
manifold applications with concerted operation of one or
even more coexisting physical phenomena in addition to
spin state switching, to yield, for example, signal enhancement, multisensing, and multifiltering effects.

Acknowledgments
On behalf of the Mainz group, I wish to extend sincere thanks to
various institutions that have continuously and generously supported our research work: Deutsche Forschungsgemeinschaft, Federal Government of Germany, European Community, University of
Mainz, University of Darmstadt, Fonds of the Chemical Industry,
Schott Foundation Mainz. Also, it is my sincere desire to thank
many persons for pleasant and successful collaboration: All members of the Mainz research team, students of master and doctoral
theses, postdoctoral research associates, friends, and companions
of the Spin Crossover Family, the Molecular Magnetism Community, and the Mssbauer Spectroscopy Community.

588

[1] L. Cambi, L. Szeg, Ber. Dtsch. Chem. Ges. (A and B Series)


1931, 64, 2591.
[2] L. Cambi, L. Szeg, Ber. Dtsch. Chem. Ges. (A and B Series)
1933, 66, 656.
[3] L. Cambi, L. Malatesta, Ber. Dtsch. Chem. Ges. (A and B
Series) 1937, 70, 2067.
[4] C. J. Ballhausen, Introduction to Ligand Field Theory, McGrawHill, New York, 1962.
[5] H. L. Schlfer, G. Gliemann, Einfhrung in die Ligandenfeldtheorie, Akademische Verlagsgesellschaft, Frankfurt/Main,
1967.
[6] A. Hauser, Top. Curr. Chem. 2004, 233, 49 (Eds.: P. Gtlich,
H. A. Goodwin).
[7] P. Gtlich, A. Hauser, H. Spiering, Angew. Chem. 1994, 106,
2109; Angew. Chem. Int. Ed. Engl. 1994, 33, 2024.
[8] J. S. Griffith, L. E. Orgel, Quart. Rev. (Chem. Soc. London)
1957, 11, 381.
[9] Y. Garcia, P. Gtlich, Top. Curr. Chem. 2004, 234, 49 (Eds.: P.
Gtlich, H. A. Goodwin).
[10] P. Gtlich, H. A. Goodwin (Eds.), Topics in Current Chemistry,
Vols. 233, 234, 235, Springer, Berlin, 2004.
[11] P. J. van Koningsbruggen, Y. Maeda, H. Oshio, Top. Curr.
Chem. 2004, 233, 259 (Eds.: P. Gtlich, H. A. Goodwin).
2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.eurjic.org
[12] B. Weber, I. Kapplinger, H. Gorls, E. G. Jger, Eur. J. Inorg.
Chem. 2005, 2794.
[13] C. Enachescu, A. Hauser, J. J. Girerd, M. L. Boillot, ChemPhysChem 2006, 7, 1127.
[14] S. Floquet, N. Guillou, P. Negrier, E. Riviere, M. L. Boillot,
New J. Chem. 2006, 30, 1621.
[15] S. Hayami, S. Miyazaki, M. Yamamoto, K. Hiki, N. Motokawa, A. Shuto, K. Inoue, T. Shinmyozu, Y. Maeda, Bull. Chem.
Soc. Jpn. 2006, 79, 442.
[16] M. Nihei, T. Shiga, Y. Maeda, H. Oshio, Coord. Chem. Rev.
2007, 251, 2606.
[17] S. V. Larionov, Russ. J. Coord. Chem. 2008, 34, 237.
[18] K. S. Murray, Eur. J. Inorg. Chem. 2008, 3101.
[19] R. Pritchard, S. A. Barrett, C. A. Kilner, M. A. Halcrow, Dalton Trans. 2008, 3159.
[20] R. Boca, I. Nemec, I. Salitros, J. Pavlik, R. Herchel, F. Renz,
Pure Appl. Chem. 2009, 81, 1357.
[21] S. Hayami, K. Hiki, T. Kawahara, Y. Maeda, D. Urakami, K.
Inoue, M. Ohama, S. Kawata, O. Sato, Chem. Eur. J. 2009, 15,
3497.
[22] S. Imatomi, S. Hashimoto, N. Matsumoto, Eur. J. Inorg. Chem.
2009, 721.
[23] H. J. Krger, Coord. Chem. Rev. 2009, 253, 2450.
[24] M. Lorenc, J. Hebert, N. Moisan, E. Trzop, M. Servol, M.
Buron-Le Cointe, H. Cailleau, M. L. Boillot, E. Pontecorvo,
M. Wulff, S. Koshihara, E. Collet, Phys. Rev. Lett. 2009, 103.
[25] I. Salitros, N. T. Madhu, R. Boca, J. Pavlik, M. Ruben, Monatsh. Chem. 2009, 140, 695.
[26] S. Hayami, D. Urakami, Y. Yamamoto, K. Kato, Y. Kojima,
S. Nakashima, K. Inoue, Chem. Lett. 2010, 39, 328.
[27] B. Weber, E. G. Jger, Eur. J. Inorg. Chem. 2009, 465.
[28] N. Kojima, M. Enomoto, N. Kida, K. Kagesawa, Materials
2010, 3, 3141.
[29] M. Clemente-Leon, E. Coronado, M. Lopez-Jorda, J. C. Waerenborgh, Inorg. Chem. 2011, 50, 9122.
[30] A. C. Bowman, C. Milsmann, E. Bill, Z. R. Turner, E. Lobkovsky, S. DeBeer, K. Wieghardt, P. J. Chirik, J. Am. Chem. Soc.
2011, 133, 17353.
[31] S. Mossin, B. L. Tran, D. Adhikari, M. Pink, F. W. Heinemann,
J. Sutter, R. K. Szilagyi, K. Meyer, D. J. Mindiola, J. Am.
Chem. Soc. 2012, 134, 13651.
[32] A. Tissot, L. Rechignat, A. Bousseksou, M. L. Boillot, J. Mater. Chem. 2012, 22, 3411.
[33] K. Madeja, E. Knig, J. Inorg. Nucl. Chem. 1963, 25, 377.
[34] W. A. Baker, H. M. Bobonich, Inorg. Chem. 1964, 3, 1184.
[35] E. Knig, K. Madeja, Chem. Commun. (London) 1966, 61.
[36] R. L. Mssbauer, Z. Phys. 1958, 151, 124.
[37] R. L. Mssbauer, Naturwissenschaften 1958, 45, 538.
[38] P. Gtlich, A. Bill, A. X. Trautwein, Mssbauer Spectroscopy
and Transition Metal Chemistry Fundamentals and Applications, Springer, Heidelberg, 2011.
[39] P. Gtlich, Z. Anorg. Allg. Chem. 2012, 638, 15.
[40] E. Knig, Coord. Chem. Rev. 1968, 3, 471.
[41] H. A. Goodwin, Coord. Chem. Rev. 1976, 18, 293.
[42] P. Gtlich, Struct. Bonding (Berlin) 1981, 44, 83.
[43] H. Toftlund, Coord. Chem. Rev. 1989, 94, 67.
[44] P. Gtlich, A. Hauser, Coord. Chem. Rev. 1990, 97, 1.
[45] P. Gtlich, Y. Garcia, H. A. Goodwin, Chem. Soc. Rev. 2000,
29, 419.
[46] J. A. Real, A. B. Gaspar, V. Niel, M. C. Muoz, Coord. Chem.
Rev. 2003, 236, 121.
[47] M. A. Halcrow, Polyhedron 2007, 26, 3523.
[48] J. A. Kitchen, S. Brooker, Coord. Chem. Rev. 2008, 252, 2072.
[49] A. B. Gaspar, M. Seredyuk, P. Gtlich, Coord. Chem. Rev.
2009, 253, 2399.
[50] M. A. Halcrow, Coord. Chem. Rev. 2009, 253, 2493.
[51] B. Weber, Coord. Chem. Rev. 2009, 253, 2432.
[52] D. Boinnard, A. Bousseksou, A. Dworkin, J. M. Savariault, F.
Varret, J. P. Tuchagues, Inorg. Chem. 1994, 33, 271.
Eur. J. Inorg. Chem. 2013, 581591

ESSAY

[53] B. Weber, E. Kaps, J. Weigand, C. Carbonera, J. F. Ltard, K.


Achterhold, F. G. Parak, Inorg. Chem. 2008, 47, 487.
[54] J. Klingele, D. Kaase, J. Hilgert, G. Steinfeld, M. H. Klingele,
J. Lach, Dalton Trans. 2010, 39, 4495.
[55] E. Knig, G. Ritter, S. K. Kulshreshtha, J. Waigel, L. Sacconi,
Inorg. Chem. 1984, 23, 12411246.
[56] H. A. Goodwin, Top. Curr. Chem. 2004, 233, 59 (Eds.: P.
Gtlich, H. A. Goodwin).
[57] G. J. Long, F. Grandjean, D. L. Reger, Top. Curr. Chem. 2004,
233, 91 (Eds.: P. Gtlich, H. A. Goodwin).
[58] P. J. van Koningsbruggen, Top. Curr. Chem. 2004, 233, 123
(Eds.: P. Gtlich, H. A. Goodwin).
[59] H. Toftlund, J. J. McGarvey, Top. Curr. Chem. 2004, 233, 151
(Eds.: P. Gtlich, H. A. Goodwin).
[60] J. A. Real, A. B. Gaspar, M. C. Munoz, P. Gtlich, V. Ksenofontov, H. Spiering, Top. Curr. Chem. 2004, 233, 167 (Eds.: P.
Gtlich, H. A. Goodwin).
[61] Y. Garcia, V. Niel, M. C. Munoz, J. A. Real, Top. Curr. Chem.
2004, 233, 229 (Eds.: P. Gtlich, H. A. Goodwin).
[62] R. C. Stoufer, D. H. Busch, W. B. Hadley, J. Am. Chem. Soc.
1961, 83, 3732.
[63] L. Sacconi, Coord. Chem. Rev. 1972, 8, 351.
[64] H. A. Goodwin, Top. Curr. Chem. 2004, 234, 23 (Eds.: P.
Gtlich, H. A. Goodwin).
[65] D. M. Jenkins, J. C. Peters, J. Am. Chem. Soc. 2005, 127, 7148.
[66] I. Krivokapic, M. Zerara, M. L. Daku, A. Vargas, C. Enachescu, C. Ambrus, P. Tregenna-Piggott, N. Amstutz, E.
Krausz, A. Hauser, Coord. Chem. Rev. 2007, 251, 364.
[67] O. Sato, A. L. Cui, R. Matsuda, J. Tao, S. Hayami, Acc. Chem.
Res. 2007, 40, 361.
[68] S. Hayami, Y. Komatsu, T. Shimizu, H. Kamihata, Y. N. Lee,
Coord. Chem. Rev. 2011, 255, 1981.
[69] G. G. Morgan, K. D. Murnaghan, H. Mller-Bunz, V. McKee,
C. J. Harding, Angew. Chem. 2006, 118, 7350; Angew. Chem.
Int. Ed. 2006, 45, 7192.
[70] Z. L. Liu, S. L. Liang, X. W. Di, J. Zhang, Inorg. Chem. Commun. 2008, 11, 783.
[71] W. Klui, J. Chem. Soc., Chem. Commun. 1979, 700.
[72] W. Klui, Z. Naturforsch. B 1979, 34, 1402.
[73] W. Klui, W. Eberspach, P. Gtlich, Inorg. Chem. 1987, 26,
3977.
[74] P. Gtlich, B. R. McGarvey, W. Klui, Inorg. Chem. 1980, 19,
3704.
[75] S. Thies, C. Bornholdt, F. Kohler, F. D. Sonnichsen, C. Nather,
F. Tuczek, R. Herges, Chem. Eur. J. 2010, 16, 10074.
[76] S. Thies, H. Sell, C. Schutt, C. Bornholdt, C. Nather, F. Tuczek,
R. Herges, J. Am. Chem. Soc. 2011, 133, 16243.
[77] S. Venkataramani, U. Jana, M. Dommaschk, F. D. Sonnichsen,
F. Tuczek, R. Herges, Science 2011, 331, 445.
[78] a) E. Gerdau, H. de Waard (Eds.), Hyperfine Interact., Vol.
123/124 1999; b) E. Gerdau, H. de Waard (Eds.), Hyperfine
Interact. 2000, 125; c) H. Winkler, A. I. Chumakov, A. X.
Trautwein, Top. Curr. Chem. 2004, 235, 137 (Eds.: P. Gtlich,
H. A. Goodwin).
[79] H. Grunsteudel, H. Paulsen, W. MeyerKlaucke, H. Winkler,
A. X. Trautwein, H. F. Grunsteudel, A. Q. R. Baron, A. I.
Chumakov, R. Rffer, H. Toftlund, Hyperfine Interact. 1998,
113, 311.
[80] L. H. Bttger, A. I. Chumakov, C. M. Grunert, P. Gutlich, J.
Kusz, H. Paulsen, U. Ponkratz, V. Rusanov, A. X. Trautwein,
J. A. Wolny, Chem. Phys. Lett. 2006, 429, 189.
[81] K. L. Ronayne, H. Paulsen, A. Hofer, A. C. Dennis, J. A.
Wolny, A. I. Chumakov, V. Schnemann, H. Winkler, H. Spiering, A. Bousseksou, P. Gtlich, A. X. Trautwein, J. J. McGarvey, Phys. Chem. Chem. Phys. 2006, 8, 4685.
[82] P. Guionneau, M. Marchivie, G. Bravic, J. F. Ltard, D. Chasseau, Top. Curr. Chem. 2004, 234, 97 (Eds.: P. Gtlich, H. A.
Goodwin).
[83] J. Kusz, P. Gtlich, H. Spiering, Top. Curr. Chem. 2004, 234,
129 (Eds.: P. Gtlich, H. A. Goodwin).
589

2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.eurjic.org
[84] E. Knig, Prog. Inorg. Chem. 1987, 35, 527.
[85] M. Sorai, J. Ensling, P. Gtlich, Chem. Phys. 1976, 18, 199.
[86] M. A. Halcrow, Chem. Soc. Rev. 2011, 40, 4119.
[87] M. Sorai, Top. Curr. Chem. 2004, 235, 153 (Eds.: P. Gtlich,
H. A. Goodwin).
[88] M. Sorai, M. Nakano, Y. Miyazaki, Chem. Rev. 2006, 106, 976.
[89] G. A. Berezovskii, L. G. Lavrenova, J. Therm. Anal. Calorim.
2011, 103, 1063.
[90] M. Sorai, S. Seki, J. Phys. Soc. Jpn. 1972, 33, 575.
[91] M. Sorai, S. Seki, J. Phys. Chem. Solids 1974, 35, 555.
[92] B. Weber, F. A. Walker, Inorg. Chem. 2007, 46, 6794.
[93] J. J. Girerd, M. L. Boillot, G. Blain, E. Riviere, Inorg. Chim.
Acta 2008, 361, 4012.
[94] I. V. Ovchinnikov, T. A. Ivanova, A. N. Turanov, R. R. Garipov, Phys. Solid State 2009, 51, 2058.
[95] J. E. Penner-Hahn, Coord. Chem. Rev. 1999, 190, 1101.
[96] A. Michalowicz, J. Moscovici, B. Ducourant, D. Cracco, O.
Kahn, Chem. Mater. 1995, 7, 1833.
[97] Y. Garcia, P. J. van Koningsbruggen, G. Bravic, P. Guionneau,
D. Chasseau, G. L. Cascarano, J. Moscovici, K. Lambert, A.
Michalowicz, O. Kahn, Inorg. Chem. 1997, 36, 6357.
[98] A. Vertes, K. Suvegh, R. Hinek, P. Gtlich, J. Phys. Chem.
Solids 1994, 55, 1269.
[99] S. J. Campbell, V. Ksenofontov, Y. Garcia, J. S. Lord, S. Reiman, P. Gtlich, Hyperfine Interact. C 2002, 5, 363.
[100] S. J. Campbell, V. Ksenofontov, Y. Garcia, J. S. Lord, Y. Boland, P. Gtlich, J. Phys. Chem. B 2003, 107, 14289.
[101] M. Sorai, J. Ensling, K. M. Hasselbach, P. Gtlich, Chem.
Phys. 1977, 20, 197.
[102] A. Bousseksou, L. Tommasi, G. Lemercier, F. Varret, J. P. Tuchagues, Chem. Phys. Lett. 1995, 243, 493.
[103] N. Willenbacher, H. Spiering, J. Phys. C Solid State Phys.
1988, 21, 1423.
[104] H. Spiering, N. Willenbacher, J. Phys. Condens. Matter 1989,
1, 10089.
[105] H. Spiering, Top. Curr. Chem. 2004, 235, 171 (Eds.: P.
Gtlich, H. A. Goodwin).
[106] H. G. Drickamer, C. W. Frank, Electronic Transitions and the
High Pressure Chemistry and Physics of Solids, Wiley, New
York, 1973.
[107] H. Drickamer, Angew. Chem. 1974, 86, 61; Angew. Chem. Int.
Ed. Engl. 1974, 13, 39.
[108] A. Bousseksou, G. Molnar, G. Matouzenko, Eur. J. Inorg.
Chem. 2004, 4353.
[109] V. Ksenofontov, A. B. Gaspar, P. Gtlich, Top. Curr. Chem.
2004, 235, 23 (Eds.: P. Gtlich, H. A. Goodwin).
[110] J. A. Real, A. B. Gaspar, M. C. Munoz, Dalton Trans. 2005,
2062.
[111] C. P. Khler, R. Jakobi, E. Meissner, L. Wiehl, H. Spiering,
P. Gtlich, J. Phys. Chem. Solids 1990, 51, 239.
[112] V. Ksenofontov, H. Spiering, A. Schreiner, G. Levchenko,
H. A. Goodwin, P. Gtlich, J. Phys. Chem. Solids 1999, 60,
393.
[113] Y. Qi, E. W. Mller, H. Spiering, P. Gtlich, Chem. Phys. Lett.
1983, 101, 503.
[114] A. Bousseksou, F. Varret, M. Goiran, K. Boukheddaden, J. P.
Tuchagues, Top. Curr. Chem. 2004, 235, 65 (Eds.: P. Gtlich,
H. A. Goodwin).
[115] a) J. K. Beattie, R. A. Binstead, R. J. West, Mol. Rate Processes, Pap. Symp., C7, 6 pp. R. Aust. Chem. Inst.: Parkville,
Aust. 1975; b) R. A. Binstead, J. K. Beattie, T. G. Dewey,
D. H. Turner, J. Am. Chem. Soc. 1980, 102, 6442; c) J. K.
Beattie, Adv. Inorg. Chem. 1988, 32, 1; d) J. K. Beattie, R. A.
Binstead, M. T. Kelso, P. DelFavero, T. G. Dewey, D. H.
Turner, Inorg. Chim. Acta 1995, 235, 245; e) M. A. Bergkamp,
B. S. Brunschwig, P. Gtlich, T. L. Netzel, N. Sutin, Chem.
Phys. Lett. 1981, 81, 147; f) C. Brady, J. J. McGarvey, J. K.
McCusker, H. Toftlund, D. N. Hendrickson, Top. Curr. Chem.
2004, 235, 1 (Eds.: P. Gtlich, H. A. Goodwin); g) J. J.
McGarvey, I. Lawthers, J. Chem. Soc., Chem. Commun. 1982,
Eur. J. Inorg. Chem. 2013, 581591

[116]
[117]

[118]
[119]
[120]
[121]

[122]

[123]

[124]

[125]
[126]
[127]
[128]
[129]
[130]
[131]

[132]
[133]

590

[134]
[135]

ESSAY

906; h) I. Lawthers, J. J. McGarvey, J. Am. Chem. Soc. 1984,


106, 4280.
P. Gtlich, Top. Curr. Chem. 2004, 234, 231 (Eds.: P. Gtlich,
H. A. Goodwin).
a) S. Decurtins, P. Gtlich, C. P. Khler, H. Spiering, A.
Hauser, Chem. Phys. Lett. 1984, 105, 1; b) S. Decurtins, P.
Gtlich, K. M. Hasselbach, A. Hauser, H. Spiering, Inorg.
Chem. 1985, 24, 2174.
a) A. Hauser, Chem. Phys. Lett. 1986, 124, 543; b) A. Hauser,
Coord. Chem. Rev. 1991, 111, 275.
A. Hauser, Top. Curr. Chem. 2004, 234, 155 (Eds.: P. Gtlich,
H. A. Goodwin).
F. Renz, H. Oshio, V. Ksenofontov, M. Waldeck, H. Spiering,
P. Gtlich, Angew. Chem. 2000, 112, 3832; Angew. Chem. Int.
Ed. 2000, 39, 3699.
a) J. F. Ltard, J. Mater. Chem. 2006, 16, 2550; b) J. F. Letard,
P. Guionneau, O. Nguyen, J. S. Costa, S. Marcen, G. Chastanet, M. Marchivie, L. Goux-Capes, Chem. Eur. J. 2005, 11,
4582.
a) G. Chastanet, C. Carbonera, C. Mingotaud, J. F. Ltard,
J. Mater. Chem. 2004, 14, 3516; b) A. Bousseksou, G. Molnar,
J. A. Real, K. Tanaka, Coord. Chem. Rev. 2007, 251, 1822; c)
J. F. Letard, C. Carbonera, J. A. Real, S. Kawata, S. Kaizaki,
Chem. Eur. J. 2009, 15, 4146; d) A. Bhattacharjee, V. Ksenofontov, J. A. Kitchen, N. G. White, S. Brooker, P. Gtlich,
Appl. Phys. Lett. 2008, 92, 174104.
a) P. Rosa, A. Debay, L. Capes, G. Chastanet, A. Bousseksou,
P. Le Floch, J. F. Ltard, Eur. J. Inorg. Chem. 2004, 3017; b)
M. Quesada, F. Prins, O. Roubeau, P. Gamez, S. J. Teat, P. J.
van Koningsbruggen, J. G. Haasnoot, J. Reedijk, Inorg. Chim.
Acta 2007, 360, 3787.
a) S. Hayami, K. Danjobara, S. Miyazaki, K. Inoue, Y.
Ogawa, Y. Maeda, Polyhedron 2005, 24, 2821; b) S. Hayami,
N. Motokawa, A. Shuto, R. Moriyama, N. Masuhara, K. Inoue, Y. Maeda, Polyhedron 2007, 26, 2375; c) S. Hayami, D.
Urakami, Y. Kojima, H. Yoshizaki, Y. Yamamoto, K. Kato,
A. Fuyuhiro, S. Kawata, K. Inoue, Inorg. Chem. 2010, 49,
1428.
F. Renz, H. Spiering, H. A. Goodwin, P. Gtlich, Hyperfine
Interact. 2000, 126, 155.
F. Varret, K. Boukheddaden, E. Codjovi, C. Enachescu, J.
Linares, Top. Curr. Chem. 2004, 234, 199 (Eds.: P. Gtlich,
H. A. Goodwin).
G. Vanko, F. Renz, G. Molnar, T. Neisius, S. Karpati, Angew.
Chem. 2007, 119, 5400; Angew. Chem. Int. Ed. 2007, 46, 5306.
F. Renz, G. Vanko, P. Homenya, R. Saadat, Z. Nemeth, S.
Huotari, Eur. J. Inorg. Chem. 2012, 2653.
T. G. Gopakumar, F. Matino, H. Naggert, A. Bannwarth, F.
Tuczek, R. Berndt, Angew. Chem. 2012, 124, 63676371; Angew. Chem. Int. Ed. 2012, 51, 6262.
M. L. Boillot, J. Zarembowitch, A. Sour, Top. Curr. Chem.
2004, 234, 261 (Eds.: P. Gtlich, H. A. Goodwin).
a) M. M. N. Wolf, R. Gross, C. Schumann, J. A. Wolny, V.
Schnemann, A. Dossing, H. Paulsen, J. J. McGarvey, R.
Diller, Phys. Chem. Chem. Phys. 2008, 10, 4264; b) J. A.
Wolny, R. Diller, V. Schnemann, Eur. J. Inorg. Chem. 2012,
2635.
T. G. Gopakumar, F. Matino, H. Naggert, A. Bannwarth, F.
Tuczek, R. Berndt, Angew. Chem. 2012, 124, 6367; Angew.
Chem. Int. Ed. 2012, 51, 6262.
H. Paulsen, A. X. Trautwein, Top. Curr. Chem. 2004, 235, 197
(Eds.: P. Gtlich, H. A. Goodwin).
A. Real, J. Zarembowitch, O. Kahn, X. Solans, Inorg. Chem.
1987, 26, 2939.
a) G. Vos, R. A. Le Febre, R. A. G. De Graaff, J. G. Haasnoot, J. Reedijk, J. Am. Chem. Soc. 1983, 105, 1682; b) P. J.
van Koningsbruggen, J. G. Haasnoot, W. Vreugdenhil, J. Reedijk, O. Kahn, Inorg. Chim. ACTA 1995, 239, 5; c) J. J. A.
Kolnaar, G. vanDijk, H. Kooijman, A. L. Spek, V. G. Ksenofontov, P. Gtlich, J. G. Haasnoot, J. Reedijk, Inorg. Chem.
2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.eurjic.org

[136]

[137]
[138]
[139]
[140]
[141]
[142]
[143]

[144]

[145]

1997, 36, 2433; d) G. Psomas, N. Brefuel, F. Dahan, J. P. Tuchagues, Inorg. Chem. 2004, 43, 4590.
a) E. Breuning, M. Ruben, J. M. Lehn, F. Renz, Y. Garcia, V.
Ksenofontov, P. Gtlich, E. Wegelius, K. Rissanen, Angew.
Chem. 2000, 112, 2563; Angew. Chem. Int. Ed. 2000, 39, 2504;
b) M. Ruben, E. Breuning, J. M. Lehn, V. Ksenofontov, F.
Renz, P. Gtlich, G. B. M. Vaughan, Chem. Eur. J. 2003, 9,
4422.
J. J. A. Kolnaar, M. I. de Heer, H. Kooijman, A. L. Spek, G.
Schmitt, V. Ksenofontov, P. Gtlich, J. G. Haasnoot, J. Reedijk, Eur. J. Inorg. Chem. 1999, 881.
J. G. Haasnoot, G. Vos, W. L. Groeneveld, Z. Naturforsch. B
1977, 32, 1421.
L. G. Lavrenova, V. N. Ikorskii, V. A. Varnek, I. M. Oglezneva, S. V. Larionov, Koord. Khim. 1986, 12, 207.
M. C. Munoz, J. A. Real, Coord. Chem. Rev. 2011, 255, 2068.
O. Kahn, C. J. Martinez, Science 1998, 279, 44.
J. F. Ltard, P. Guionneau, L. Goux-Capes, Top. Curr. Chem.
2004, 235, 221 (Eds.: P. Gtlich, H. A. Goodwin).
a) D. G. Kurth, Molecular Electronics II, Max Planck Inst.
Colloids & Interfaces, Potsdam, Germany, 2002, vol. 960, p.
29; b) S. Cobo, G. Molnar, J. A. Real, A. Bousseksou, Angew.
Chem. 2006, 118, 5918; Angew. Chem. Int. Ed. 2006, 45, 5786;
c) Y. Einaga, M. Kotake, Y. Yamada, O. Sato, Chem. Lett.
2003, 32, 846; d) O. Sato, J. Photochem. Photobiol. C: Photochem. Rev. 2004, 5, 203; e) Y. Bodenthin, U. Pietsch, H.
Mhwald, D. G. Kurth, J. Am. Chem. Soc. 2005, 127, 3110;
f) A. D. Naik, M. M. Dirtu, Y. Garcia in International Conference on the Applications of the Mossbauer Effect, Vol. 217
(Eds.: H. Mller, M. Reissner, W. Steiner, G. Wiesinger), Iop
Publishing Ltd, Bristol, 2010; g) M. Rubio, R. Hernandez, A.
Nogales, A. Roig, D. Lopez, Eur. Polym. J. 2011, 47, 52.
a) O. Roubeau, A. Colin, W. Schmitt, R. Clerac, Angew.
Chem. 2004, 116, 3345; Angew. Chem. Int. Ed. 2004, 43, 3283;
b) K. Kuroiwa, T. Shibata, S. Sasaki, M. Ohba, A. Takahara,
T. Kunitake, N. Kimizuka, J. Polym. Sci., Part A Polym.
Chem. 2006, 44, 5192; c) H. Matsukizono, K. Kuroiwa, N.
Kimizuka, J. Am. Chem. Soc. 2008, 130, 5622; d) M. Rubio,
D. Lopez, Eur. Polym. J. 2009, 45, 3339; e) P. Grondin, O.
Roubeau, M. Castro, H. Saadaoui, A. Colin, R. Clerac,
Langmuir 2010, 26, 5184.
a) C. Thibault, G. Molnar, L. Salmon, A. Bousseksou, C.
Vieu, Langmuir 2010, 26, 1557; b) M. Cavallini, I. Bergenti,
S. Milita, J. C. Kengne, D. Gentili, G. Ruani, I. Salitros, V.

Eur. J. Inorg. Chem. 2013, 581591

[146]

[147]
[148]
[149]

[150]

[151]

591

ESSAY

Meded, M. Ruben, Langmuir 2011, 27, 4076; c) C. Faulmann,


J. Chahine, I. Malfant, D. de Caro, B. Cormary, L. Valade,
Dalton Trans. 2011, 40, 2480; d) V. Martinez, I. Boldog, A. B.
Gaspar, V. Ksenofontov, A. Bhattacharjee, P. Gutlich, J. A.
Real, Chem. Mater. 2010, 22, 4271; e) J. R. Galan-Mascaros,
E. Coronado, A. Forment-Aliaga, M. Monrabal-Capilla, E.
Pinilla-Cienfuegos, M. Ceolin, Inorg. Chem. 2010, 49, 5706;
f) I. Boldog, A. B. Gaspar, V. Martinez, P. Pardo-Ibanez, V.
Ksenofontov, A. Bhattacharjee, P. Gutlich, J. A. Real, Angew.
Chem. 2008, 120, 6533; Angew. Chem. Int. Ed. 2008, 47, 6433.
a) S. Hayami, K. Kawamura, G. Juhasz, T. Kawahara, K.
Uehashi, Y. Maeda, Molecular Crystals and Liquid Crystals,
Fukuoka 8128581, Japan Kyushu Univ, Fac Sci, Dept Chem,
Higashi Ku, Fukuoka 8128581, Japan 2002, vol. 379, p. 371;
b) S. Hayami, K. Hashiguchi, G. Juhasz, M. Ohba, H. Okawa, Y. Maeda, K. Kato, K. Osaka, M. Takata, K. Inoue,
Inorg. Chem. 2004, 43, 4124.
a) K. S. Murray, C. J. Kepert, Top. Curr. Chem. 2004, 233,
195 (Eds.: P. Gtlich, H. A. Goodwin); b) S. Horike, S. Shimomura, S. Kitagawa, Nat. Chem. 2009, 1, 695.
A. B. Gaspar, V. Ksenofontov, M. Seredyuk, P. Gtlich, Coord. Chem. Rev. 2005, 249, 2661.
a) S. Hayami, K. Danjobara, K. Inoue, Y. Ogawa, N. Matsumoto, Y. Maeda, Adv. Mater. 2004, 16, 869; b) S. Hayami, N.
Motokawa, A. Shuto, N. Masuhara, T. Someya, Y. Ogawa,
K. Inoue, Y. Maeda, Inorg. Chem. 2007, 46, 1789; c) S. Hayami, R. Moriyama, A. Shuto, Y. Maeda, K. Ohta, K. Inoue,
Inorg. Chem. 2007, 46, 7692.
a) C. Faulmann, S. Dorbes, W. G. de Bonneval, G. Molnar,
A. Bousseksou, C. J. Gomez-Garcia, E. Coronado, L. Valade,
Eur. J. Inorg. Chem. 2005, 3261; b) M. Nihei, Y. Sekine, N.
Suganami, K. Nakazawa, A. Nakao, H. Nakao, Y. Murakami, H. Oshio, J. Am. Chem. Soc. 2011, 133, 3592; c) M.
Nihei, N. Takahashi, H. Nishikawa, H. Oshio, Dalton Trans.
2011, 40, 2154; d) T. Mahfoud, G. Molnar, S. Cobo, L.
Salmon, C. Thibault, C. Vieu, P. Demont, A. Bousseksou,
Appl. Phys. Lett. 2011, 99; e) S. Dorbes, L. Valade, J. A. Real,
C. Faulmann, Chem. Commun. 2005, 69; f) C. Faulmann, K.
Jacob, S. Dorbes, S. Lampert, I. Malfant, M. L. Doublet, L.
Valade, J. A. Real, Inorg. Chem. 2007, 46, 8548.
L. Salmon, G. Molnar, D. Zitouni, C. Quintero, C. Bergaud,
J. C. Micheau, A. Bousseksou, J. Mater. Chem. 2010, 20,
5499.
Received: January 20, 2013

2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

S-ar putea să vă placă și