Vous êtes sur la page 1sur 25

Int. J. Nanotechnology, Vol. x, No.

x, xxxx

Molecular Nanomagnets: towards molecular


spintronics
Wolfgang Wernsdorfer
Institut Neel, CNRS & Universite J. Fourier, BP 166,
25 rue des Martyrs, 38042 GRENOBLE Cedex 9, France
Fax: + 33 476 88 1191
E-mail: wolfgang.wernsdorfer@grenoble.cnrs.fr
Abstract: Molecular nanomagnets, often called single-molecule magnets, have
attracted much interest in recent years both from experimental and theoretical
point of view. These systems are organometallic clusters characterized by a large
spin ground state with a predominant uniaxial anisotropy. The quantum nature of
these systems makes them very appealing for phenomena occurring on the mesoscopic scale, i.e., at the boundary between classical and quantum physics. Below
their blocking temperature, they exhibit magnetization hysteresis, the classical
macroscale property of a magnet, as well as quantum tunneling of magnetization
and quantum phase interference, the properties of a microscale entity. Quantum effects are advantageous for some potential applications of single-molecule
magnets, e.g. in providing the quantum superposition of states for quantum computing, but are a disadvantage in others such as information storage. It is believed
that single-molecule magnets have a potential for quantum computation, in particular because they are extremely small and almost identical, allowing to obtain,
in a single measurement, statistical averages of a larger number of qubits. This
review introduces few basic concepts that are needed to understand the quantum
phenomena observed in molecular nanomagnets and discusses new trends of the
field of molecular nanomagnets towards molecular spintronics.
Keywords: Single-molecule magnets, molecular nanomagnets, molecular spintronics, magnetic hysteresis, resonant quantum tunneling, quantum interference,
spin parity effect, decoherence, quantum computation, qubit, exchange-bias,
spin-Hamiltonian, micro-SQUID, magnetometer.
Biographical notes: Dr. Wolfgang Wernsdorfer, born in Wurzburg, Germany, in 1966, received his education in Physics in Wurzburg, Lyon, and then
Grenoble, where he is at present Research Director at the Centre National de
la Recherche Scientifique. During his PhD in the low-temperature laboratory
(CNRS, Grenoble) Wolfgang Wernsdorfer and collaborators developed a unique
device (micro-SQUID) for measuring magnetic properties of nanostructures with
a billion times higher sensitivity than commercial magnetometers (Bronze Medal
from CNRS, 1998). His instrument allows observation of the magnetic behavior
of nanomagnets containing less than a thousand magnetic centers, which is still
a world record. Using the unique advantages of this device, Wolfgang Wernsdorfer has studied a variety of peculiar phenomena in depth, such as tunnelling
of magnetization in molecular clusters, leading to the Agilent Europhysics Prize
in 2002 and the International Olivier Kahn Award in 2006.
Over the years, the innovative approach to such studies combined with the recognized superiority of this micro-SQUID have led to worldwide collaboration
with most other notorious groups working on synthesizing molecular magnets
to investigate single-molecule magnet behavior in more than 350 systems. The

c 200x Inderscience Enterprises Ltd.


Copyright

Wolfgang Wernsdorfer
leading work of Wolfgang Wernsdorfer and collaborators in this field is at the
heart of todays knowledge on molecular magnetism.

1 Introduction
A revolution in electronics is in view, with the contemporary evolution of two novel
disciplines, spintronics and molecular electronics. A link between these two fields can be
established using molecular magnetic materials and, in particular, single-molecule magnets, which combine the classic macroscale properties of a magnet with the quantum properties of a nanoscale entity. The resulting field, molecular spintronics aims at manipulating
spins and charges in electronic devices containing one or more molecules [1, 2, 3].
The contemporary exploitation of electronic charge and spin degrees of freedom is a
particularly promising field both at fundamental and applied levels. This discipline, called
spintronics, has already seen some of its fundamental results turned into actual devices in a
record time of 10 years and it holds great promises for the future [4, 5]. Spintronic systems
exploit the fact that the electron current is composed of spin-up and spin-down carriers
that carry information encoded in their spin state and interact with magnetic materials
differently. Information encoded in spins persists when the device is switched off; it can be
manipulated with and without using magnetic fields and can be written using little energy,
to cite just a few advantages of this approach.
New efforts are now directed towards spintronic devices that preserve and exploit quantum coherence, so that fundamental investigations are shifting from metals to semiconducting [4, 5], and organic materials [6], which potentially offer best promises for cost, integration and versatility. For example, organic materials are already used in applications such
as organic light-emitting diodes (OLED), displays and organic transistors. The concomitant trend towards ever-smaller electronic devices (having already reached the nano-scale),
and the tailoring of new molecules possessing increased conductance and functionalities
are driving electronics to its ultimate molecular-scale limit [7], and the so-called molecular
electronics is now being intensively investigated.
In experiments of molecular electronics, the measuring devices are usually constituted
by two nanoelectrodes and a bridging molecule in between, allowing the measurement
of electron transport through single molecules. As the measurement is performed at the
molecular level, the observables are connected to molecular orbitals and not to Bloch
waves as in bulk materials. Hence, new rules are found for these systems and it becomes
possible to probe the quantum properties of the molecule directly. The electron tunnelling
processes in the electrode-molecule-electrode system can show the presence of Kondo or
Coulomb-blockade effects, depending on the binding strength between the molecule and
the electrodes, which can be tuned by selecting the appropriate chemical functional groups.
In this context, a new field of molecular spintronics is emerging that combines the
concepts and the advantages of spintronics and molecular electronics [1, 8] which requires
the creation of molecular devices using one or few magnetic molecules. Compounds of the
Single-Molecule Magnets (SMMs) class seem particularly attractive: their magnetization
relaxation time is extremely long at low temperature reaching years below 2 K with record
anisotropy barriers approaching 100 K [9]. These systems, combining the advantages of
molecular scale with the properties of bulk magnetic materials, look attractive for highdensity information storage and also, owing to their long coherence times [10, 11, 12],

Molecular spintronics

Figure 1
Representative examples of the peripheral functionalization of the outer organic shell of
the Mn12 SMM. Different functionalizations used to graft the SMM to surfaces are displayed [1, 3].
All structures are determined by X-ray crystallography, except d, which is a model structure. Solvent
molecules have been omitted. The atom color code is reported in the figure, as well as the diameter
of the clusters.

for quantum computing [13, 14, 15]. Moreover their molecular nature leads to appealing
quantum effects of the static and dynamic magnetic properties. The rich physics behind
the magnetic behaviour produces interesting effects like negative differential conductance
and complete current suppression [16, 17], which could be used in electronics. Another
advantage is that the weak spin-orbit and hyperfine interactions in organic molecules is
likely to preserve spin-coherence over time and distance much longer than in conventional
metals or semiconductors. Last but not least, specific functions (e.g. switchability with
light, electric field etc.) could be directly integrated into the molecule.
SMMs possess the right chemical characteristics to overcome several problems associated to molecular junctions. They are constituted by an inner magnetic core with a surrounding shell of organic ligands [18] that can be tailored to bind them on surfaces or into
junctions [19, 20, 21, 22] (Fig. 1). In order to strengthen magnetic interactions between
the magnetic core ions, SMMs often have delocalized bonds, which can enhance their conducting properties. SMMs come in a variety of shapes and sizes and permit selective substitutions of the ligands in order to alter the coupling to the environment [18, 19, 20, 23].
It is also possible to exchange the magnetic ions, thus changing the magnetic properties
without modifying the structure and the coupling to the environment [24, 25]. While grafting SMMs on surfaces has already led to important results, even more spectacular results
will emerge from the rational design and tuning of single SMM-based junctions.
From a physics viewpoint, SMMs are the final point in the series of smaller and smaller
units from bulk matter to atoms (Figure 2). They combine the classic macroscale properties of a magnet with the quantum properties of a nanoscale entity. They have crucial
advantages over magnetic nanoparticles in that they are perfectly monodisperse and can be

Wolfgang Wernsdorfer

Figure 2
Scale of size that goes from macroscopic down to nanoscopic sizes. The unit of this
scale is the number of magnetic moments in a magnetic system (roughly corresponding to the number
of magnetic atoms). At macroscopic sizes, a magnetic system is described by magnetic domains
that are separated by domain walls. Magnetization reversal occurs via nucleation, propagation, and
annihilation of domain walls (hysteresis loop on the left). When the system size is of the order of
magnitude of the domain wall width or the exchange length, the formation of domain walls requires
too much energy. Therefore, the magnetization remains in the so-called single-domain state, and the
magnetization reverse by uniform rotation or nonuniform modes (middle). SMMs are the final point
in the series of smaller and smaller units from bulk matter to atoms and magnetization reverses via
quantum tunneling (right).
Mesoscopic physics
Nanoscopic

Macroscopic
Permanent
magnets

Micron
particles

S = 102 0

101 0

Nanoparticles

108

106

105

Clusters

104

Molecular
clusters

103

Individual
spins

102

10

Multi-domain

Single-domain

Magnetic moments

Nucleation, propagation and


annihilation of domain walls

Uniform rotation
Curling

Resonant tunneling, quantization,


quantum thermodynamics

1
0.7K

M / MS

M / MS

M / MS

Fe8
0

0
1K

-1
-40

-1
-20

0
20
0 H(mT)

40

0.1K

-1
-100

0
100
0 H(mT)

-1

0
0H ( T )

studied in molecular crystals. They display an impressive array of quantum effects (that
are observable up to higher and higher temperatures due to progress in molecular designs),
ranging from quantum tunnelling of magnetization [26, 27, 28, 29] to Berry phase interference [30, 31] and quantum coherence [10, 11, 12] with important consequences on the
physics of spintronic devices. Although the magnetic properties of SMMs can be affected
when they are deposited on surfaces or between leads [23], these systems remain a step
ahead of non-molecular nanoparticles, which show large size and anisotropy distributions,
for a low structure versatility.
This review introduces the basic concepts that are needed to understand the quantum
phenomena observed in molecular nanomagnets and shows the new trends towards molecular spintronics [1] using junctions [3] and nano-SQUIDs [2].

2 Overview of molecular nanomagnets


Molecular nanomagnets or single-molecule magnets (SMMs) are mainly organic molecules that have one or several metal centers with unpaired electrons. These polynuclear
metal complexes are surrounded by bulky ligands (often organic carboxylate ligands). The
most prominent examples are a dodecanuclear mixed-valence manganese-oxo cluster with
acetate ligands, short Mn12 acetate [32], and an octanuclear iron(III) oxo-hydroxo cluster
of formula [Fe8 O2 (OH)12 (tacn)6 ]8+ where tacn is a macrocyclic ligand, short Fe8 [33].
Both systems have a spin ground state of S = 10 and an Ising-type magnetic anisotropy,
which stabilizes the spin states with m = 10 and generates an energy barrier for the

Molecular spintronics

Figure 3
Size scale spanning atomic to nanoscale dimensions. On the far right is shown a highresolution transmission electron microscopyview along a [110] direction of a typical 3 nm diameter
cobalt nanoparticle exhibiting a face-centered cubic structure and containing about 1000 Co atoms.
The Mn84 molecule is a 4.2 nm diameter particle. Also shown for comparison are the indicated
smaller Mn nanomagnets, which are drawn to scale. An alternative means of comparison is the Neel
vector (N), which is the scale shown. The green arrows indicate the magnitude of the Neel vectors for
the indicated SMMs, which are 7.5, 22, 61, and 168 for Mn4 , Mn12 , Mn30 and Mn84 , respectively.

Mn4

Mn12

Mn30

Mn84

N
1

10

Quantum world
Molecular (bottom-up) approach

100

1000

Classical world
Classical (top-down) approach

reversal of the magnetization of about 67 K for Mn12 acetate [34, 35, 36] and 25 K for
Fe8 [37].
Thermally activated quantum tunneling of the magnetization has first been evidenced
in both systems [26, 27, 28, 38, 39]. Theoretical discussion of this assumes that thermal
processes (principally phonons) promote the molecules up to high levels with small quantum numbers |m|, not far below the top of the energy barrier, and the molecules then
tunnel inelastically to the other [40, 41, 42, 43, 44, 45, 46, 47]. Thus the transition
is almost entirely accomplished via thermal transitions and the characteristic relaxation
time is strongly temperature-dependent. For Fe8 , however, the relaxation time becomes
temperature-independent below 0.36 K [28, 48] showing that a pure tunneling mechanism
between the only populated ground states m = S = 10 is responsible for the relaxation of the magnetization. On the other hand in the Mn12 acetate system one sees
temperature independent relaxation only for strong applied fields and below about 0.6 K
[49, 50]. During the last years, many new molecular nanomagnets were presented (see,
for instance, Refs. [51, 52, 53, 54]) which show also tunneling at low temperatures. The
largest molecular nanomagnets is currently a Mn84 molecule [55] that has a size of a magentic nanoparticle (Figure 3). The record anisotropy barriers of 89 K is currently a Mn6
SMM [9].

3 Giant spin model for nanomagnets


A magnetic molecule, that behaves like a small nanomagnet, must have a large uniaxial
easy axis type magnetic anisotropy and a large ground state spin. A typical example is the
octanuclear iron(III) oxo-hydroxo cluster of formula [Fe8 O2 (OH)12 (tacn)6 ]8+ where tacn
is a macrocyclic ligand (1,4,7-traiazcyclononane), short Fe8 (Figure 4) [33].
The internal iron(III) ions are octahedrally coordinated to the two oxides and to four
hydroxo bridges. The outer iron(III) ions coordinate three nitrogens and three hydroxyls.

Wolfgang Wernsdorfer

Figure 4
Schematic view of the magnetic core of the Fe8 cluster. The oxygen atoms are black,
the nitrogen atoms are gray, and carbon atoms are white. The arrows represent the spin structure of
the ground state S = 10.

Spin polarized neutron scattering showed that all Fe ions have a spin 5/2, six spins up and
two down [56]. This rationalizes the S = 10 spin ground state that is in agreement with
magnetization measurements.
In principle, a multi-spin Hamiltonian can be derived taking into account of all exchange interactions and the single-ion magnetic anisotropies. However, the Hilbert space
is very large (68 106 ) and the exchange coupling constants are not well known. A giant
spin model is therefore often used that describes in an effective way the ground spin state
multiplet. A nanomagnet like the Fe8 molecular cluster has the following Hamiltonian

~ H
~
H = DSz2 + E Sx2 Sy2 + gB 0 S
(1)
Sx , Sy , and Sz are the three components of the spin operator, D and E are the anisotropy
constants which were determined via high frequency electron paramgnetic resonance (HFEPR) (D/kB 0.275 K and E/kB 0.046 K [37]), and the last term of the Hamiltonian
~ This Hamiltonian dedescribes the Zeeman energy associated with an applied field H.
fines hard, medium, and easy axes of magnetization in x, y, and z directions, respectively
(Figure 5). It has an energy level spectrum with (2S + 1) = 21 values which, to a first
approximation, can be labeled by the quantum numbers m = 10, 9, ..., 10 choosing
the z-axis as quantization axis. The energy spectrum, shown in Figure 6, can be obtained
by using standard diagonalisation techniques of the [21 21] matrix describing the spin
~ = 0, the levels m = 10 have the lowest energy. When a field
Hamiltonian S = 10. At H
Hz is applied, the energy levels with m < 2 increase, while those with m > 2 decrease
(Figure 6). Therefore, energy levels of positive and negative quantum numbers cross at
certain fields Hz . It turns out that for Fe8 the levels cross at fields given by 0 Hz n
0.22 T, with n = 1, 2, 3, .... The inset of Figure 6 displays the details at a level crossing where transverse terms containing Sx or Sy spin operators turn the crossing into an
avoided level crossing. The spin S is in resonance between two states when the local
longitudinal field is close to an avoided level crossing. The energy gap, the so-called tunnel spitting , can be tuned by an applied field in the xy-plane (Figure 5) via the Sx Hx
and Sy Hy Zeeman terms (Section 3.2).

Molecular spintronics

Figure 5
Unit sphere showing degenerate minima A and B which are joined by two tunnel paths
(heavy lines). The hard, medium, and easy axes are taken in x-, y-, and z-direction, respectively. The
constant transverse field Htrans for tunnel splitting measurements is applied in the xy-plane at an
~ = 0, the giant spin reversal results from the interference of
azimuth angle . At zero applied field H
two quantum spin paths of opposite direction in the easy anisotropy yz-plane. For transverse fields in
direction of the hard axis, the two quantum spin paths are in a plane which is parallel to the yz-plane,
as indicated in the figure. Using Stokes theorem, it has been shown that the path integrals can be
converted in an area integral, yielding that destructive interferencethat is a quench of the tunneling
rateoccurs whenever the shaded area is k/S, where k is an odd integer. The interference effects
disappear quickly when the transverse field has a component in the y-direction because the tunneling
is then dominated by only one quantum spin path.
Z

Easy axis

Hard
axis

Medium
axis

Htrans
B

Figure 6
Zeeman diagram of the 21 levels of the S = 10 manifold of Fe8 as a function of the
field applied along the easy axis [equation (1)]. From bottom to top, the levels are labeled with
quantum numbers m = 10, 9, ..., 0. The levels cross at fields given by 0 Hz n 0.22 T, with
n = 1, 2, 3, .... The inset displays the detail at a level crossing where the transverse terms (terms
containing Sx or/and Sy spin operators) turn the crossing into an avoided level crossing. The larger
the tunnel splitting , the higher the tunnel rate.

The effect of these avoided level crossings can be seen in hysteresis loop measurements
(Figure 7). When the applied field is near an avoided level crossing, the magnetization
relaxes faster, yielding steps separated by plateaus. As the temperature is lowered, there is
a decrease in the transition rate due to reduced thermal-assisted tunneling.

Wolfgang Wernsdorfer

Figure 7
Hysteresis loops of a single crystal of Fe8 molecular clusters at different temperatures.
The longitudinal field (zdirection) was swept at a constant sweeping rate of 0.014 T/s. The loops
display a series of steps, separated by plateaux. As the temperature is lowered, there is a decrease in
the transition rate due to reduced thermal assisted tunneling. The hysteresis loops become temperature independent below 0.35 K, demonstrating quantum tunneling at the lowest energy levels
1
0.7K

M / MS

0.5

0.5K
1K

-0.5
0.4, 0.3
and 0.04K

-1
-1.2

-0.6

0
0 Hz ( T )

0.6

1.2

3.1 LandauZener tunneling in Fe8


The nonadiabatic transition between the two states in a two-level system has first
been discussed by Landau, Zener, and Stuckelberg [57, 58, 59]. The original work by
Zener concentrates on the electronic states of a bi-atomic molecule, while Landau and
Stuckelberg considered two atoms that undergo a scattering process. Their solution of
the time-dependent Schrodinger equation of a two-level system could be applied to many
physical systems and it became an important tool for studying tunneling transitions. The
LandauZener model has also been applied to spin tunneling in nanoparticles and clusters
[60, 61, 62, 63, 64]. The tunneling probability P when sweeping the longitudinal field Hz
at a constant rate over an avoided energy level crossing (Figure 8) is given by
#
"
2m,m0
Pm,m0 = 1 exp
.
(2)
2~gB |m m0 |0 dHz /dt
Here, m and m0 are the quantum numbers of the avoided level crossing, dHz /dt is the
constant field sweeping rates, g 2, B the Bohr magneton, and ~ is Plancks constant.
With the LandauZener model in mind, we can now start to understand qualitatively
the hysteresis loops (Figure 7). Let us start at a large negative magnetic field Hz . At very
low temperature, all molecules are in the m = 10 ground state (Figure 6). When the
applied field Hz is ramped down to zero, all molecules will stay in the m = 10 ground
state. When ramping the field over the 10,10 region at Hz 0, there is a Landau
Zener tunnel probability P10,10 to tunnel from the m = 10 to the m = 10 state.
P10,10 depends on the sweeping rate [equation (2)]; that is, the slower the sweeping
rate, the larger the value of P10,10 . This is clearly demonstrated in the hysteresis loop
measurements showing larger steps for slower sweeping rates [30, 31]. When the field
Hz is now increased further, there is a remaining fraction of molecules in the m = 10
state which became a metastable state. The next chance to escape from this state is when
the field reaches the 10,9 region. There is a LandauZener tunnel probability P10,9 to
tunnel from the m = 10 to the m = 9 state. As m = 9 is an excited state, the molecules
in this state relax quickly to the m = 10 state by emitting a phonon. A similar mechanism
happens when the applied field reaches the 10,10n regions (n = 2, 3, . . . ) until all

Molecular spintronics

Figure 8
Detail of the energy level diagram near an avoided level crossing. m and m0 are the
quantum numbers of the energy level. Pm,m0 is the LandauZener tunnel probability when sweeping
the applied field from the left to the right over the anticrossing. The greater the gap and the slower
the sweeping rate, the higher is the tunnel rate [equation (2)].
|m>

| m' >

Energy

1 - P

P
|m>

| m' >

Magnetic field H z

molecules are in the m = 10 ground state; that is, the magnetization of all molecules is
reversed. As phonon emission can only change the molecule state by m = 1 or 2, there
is a phonon cascade for higher applied fields.
In order to apply quantitatively the LandauZener formula [equation (2)], we first saturated the crystal of Fe8 clusters in a field of Hz = 1.4 T, yielding an initial magnetization
Min = Ms . Then, we swept the applied field at a constant rate over one of the resonance
transitions and measured the fraction of molecules which reversed their spin. This procedure yields the tunneling rate P10,10n and thus the tunnel splitting 10,10n [equation (2)] with n = 0, 1, 2, . . . .
We first checked the predicted LandauZener sweeping field dependence of the tunneling rate. We found a good agreement for sweeping rates between 10 and 0.001 T/s [30].
The deviations at lower sweeping rates are mainly due to the hole-digging mechanism [65]
which slows down the relaxation. Our measurements showed for the first time that the
LandauZener method is particularly adapted for molecular clusters because it works even
in the presence of dipolar fields which spread the resonance transition provided that the
field sweeping rate is not too small.
3.2 Oscillations of tunnel splitting
An applied field in the xyplane adjusts the tunnel splittings m,m0 via the Sx and Sy
spin operators of the Zeeman terms that do not commute with the spin Hamiltonian. This
effect can be demonstrated by using the LandauZener method (Section 3.1). Figure 9
presents a detailed study of the tunnel splitting 10 at the tunnel transition between m =
10, as a function of transverse fields applied at different angles , defined as the azimuth
angle between the anisotropy hard axis and the transverse field (Figure 5). For small angles
the tunneling rate oscillates with a period of 0.4 T, whereas no oscillations showed
up for large angles [30]. In the latter case, a much stronger increase of 10 with
transverse field is observed. The transverse field dependence of the tunneling rate for
different resonance conditions between the state m = 10 and (10 n) can be observed
by sweeping the longitudinal field around 0 Hz = n 0.22 T with n = 0, 1, 2, . . . .
The corresponding tunnel splittings 10,10n oscillate with almost the same period of
0.4 T (Figure 9). In addition, comparing quantum transitions between m = 10 and
(10 n), with n even or odd, revealed a parity (or symmetry) effect that is analogous to

10

Wolfgang Wernsdorfer

Tunnel splitting (10-

50

(a)
20
7

0.1
0

(b)

K)

90

10

Tunnel splitting (10-

K)

Figure 9
Measured tunnel splitting as a function of transverse field for (a) several azimuth
angles at m = 10 and (b) 0 , as well as for quantum transition between m = 10 and
(10 n). Note the parity effect that is analogous to the suppression of tunneling predicted for halfinteger spins. It should also be mentioned that internal dipolar and hyperfine fields hinder a quench
of which is predicted for an isolated spin.

0
n=0
0.2
0.4
0.6
0.8
1
1.2
Magnetic transverse field (T)

1.4

n=2
10

0
n=1

n=0

0.1
-0.4 -0.2 0 0.2 0.4 0.6 0.8 1 1.2 1.4
Magnetic transverse field (T)

the Kramers suppression of tunneling predicted for half-integer spins [66, 67]. A similar
strong dependence on the azimuth angle was observed for all studied resonances.
3.3 Semiclassical descriptions
Before showing that the above results can be derived by an exact numerical calculation
using the quantum operator formalism, it is useful to discuss semiclassical models. The
original prediction of oscillation of the tunnel splitting was done by using the path integral
formalism [68]. Here [69], the oscillations are explained by constructive or destructive
interference of quantum spin phases (Berry phases) of two tunnel paths (instanton trajectories) (Figure 5). Since our experiments were reported, the WentzelKramersBrillouin
theory has been used independently by Garg [70] and Villain and Fort [71]. The surprise
is that although these models [69, 70, 71] are derived semiclassically, and should have
higher-order corrections in 1/S, they appear to be exact as written! This has first been
noted in Refs. [70] and [71] and then proven in Ref. [72] Some extensions or alternative
explications of Gargs result can be found in Refs. [73, 74, 75, 76, 77].
The period of oscillation is given by [69]
2kB p
H =
2E(E + D)
(3)
gB
where D and E are defined in equation (1). We find a period of oscillation of H = 0.26 T
for D = 0.275 K and E = 0.046 K as in Ref. [37]. This is somewhat smaller than the
experimental value of 0.4 T. We believe that this is due to higher-order terms of the spin
Hamiltonian which are neglected in Gargs calculation. These terms can easily be included
in the operator formalism as shown in the next subsection.
3.4 Exact numerical diagonalization
In order to quantitatively reproduce the observed periodicity we included fourth-order
terms in the spin Hamiltonian [equation (1)] as employed in the simulation of inelastic
neutron scattering measurements [78, 79] and performed a diagonalization of the [21 21]

Molecular spintronics

11

Figure 10
Calculated tunnel splitting as a function of transverse field for (a) quantum transition between m = 10 at several azimuth angles and (b) quantum transition between m = 10
and (10 n) at = 0 (Section 3.4). The fourth-order terms suppress the oscillations of at large
transverse fields |Hx |.
1000

1000
30
20

10

10
5

1
0

0.1

0.01
0

(b)

K)

(a)

100

50

Tunnel splitting (10-

Tunnel splitting (10-

K)

= 90

0.4
0.8
1.2
1.6
Magnetic tranverse field (T)

100
10
1

n=2
n=1
n=0

0.1

0.01
0

0.4
0.8
1.2
1.6
Magnetic tranverse field (T)

matrix describing the S = 10 system. For the calculation of the tunnel splitting we used
D = 0.289 K, E = 0.055 K [equation (1)] and the fourth-order terms as defined in [78]
with B40 = 0.72 106 K, B42 = 1.01 105 K, B44 = 0.43 104 K, which are close
to the values obtained by EPR measurements [80] and neutron scattering measurements
[79].
The calculated tunnel splittings for the states involved in the tunneling process at the
resonances n = 0, 1, and 2 are reported in Figure 10, showing the oscillations as well as
the parity effect for odd resonances.
3.5 Spin-parity effect
The spin-parity effect is among the most interesting quantum phenomena that can be
studied at the mesoscopic level in SMMs. It predicts that quantum tunneling is suppressed
at zero applied field if the total spin of the magnetic system is half-integer but is allowed
in integer spin systems. Enz, Schilling, Van Hemmen and Suto [81, 82] were the first to
suggest the absence of tunneling as a consequence of Kramers degeneracy. The Kramers
theorem asserts that no matter how unsymmetric the crystal field is, an ion possessing an
odd number of electrons must have a ground state which is at least doubly degenerate, even
in the presence of crystal fields and spin-orbit interactions [83].
The predicted spin parity effect can be observed by measuring the tunnel splitting as
a function of transverse field [84]. An integer spin system is rather insensitive to small
transverse fields whereas a half-integer spin systems is much more sensitive. However,
a half-integer spin system will also undergo tunneling at zero external field as a result
of environmental degrees of freedom such as hyperfine and dipolar couplings or small
intermolecular exchange interaction.
The nicest observation of the spin parity effect has been seen for two molecular Mn12
clusters with a spin ground state of S = 10 and S = 19/2 showing oscillations of the
tunnel probability as a function of a transverse field being due to topological quantum
phase interference of two tunnel paths of opposite windings (Section 3.3). Spin-parity
dependent tunneling was established for the first time in these compounds by comparing
the quantum phase interference of integer and half-integer spin systems [31].

12

Wolfgang Wernsdorfer

Figure 11
Transport experiments on SMMs. a) Schematic using a STM tip to perform transport on surface grafted SMMs. b) Schematic of SMM-based molecular transistors, in which a
gate voltage can modulate transport. c) [Co(TerPy)2 ] molecular magnet with alkyl spacers, permitting transport in the weakly coupled regime [85]. d) [Co(TerPy)2 ] molecular magnet with no
spacers, showing strong coupling and the Kondo effect [85]. e) Divanadium [(N,N,N-trimethyl1,4,7-triazacyclononane)2 V2 (CN)4 (-C4 N4 )] molecular magnet showing the Kondo effect only in the
charged state [86]. The color code is the same as in Fig. 1, except for Co atoms (green) and V atoms
(Orange).

4 Molecular spintronics using single-molecule magnets


Molecular spintronics combines the ideas of three novel disciplines, spintronics, molecular electronics, and quantum computing. The resulting field aims at manipulating spins
and charges in electronic devices containing one or more molecules [1]. The main advantage is that the weak spin-orbit and hyperfine interactions in organic molecules is likely to
preserve spin-coherence over time and distance much longer than in conventional metals
or semiconductors. In addition, specific functions (e.g. switchability with light, electric
field etc.) could be directly integrated into the molecule.
In order to lay the foundation of molecular spintronics, several molecular devices have
been proposed [1]: molecular spin-transistor, molecular spin-valve and spin filter, molecular double-dot devices, and carbon nanotube-based nano-SQUIDs [2]. The main purpose
is to fully control the initialization, the manipulation and the read-out of the spin states of
the molecule and to perform basic quantum operations. The main targets for the coming
years concern fundamental science as many issues, experimental, technological and theoretical, must be addressed before applications, for instance in quantum electronics, can be
realistically considered.
4.1 Molecular spin-transistor
The first scheme we consider is a magnetic molecule attached between two non-magnetic
electrodes. One possibility is to use a scanning tunneling microscope tip as the first electrode and the conducting substrate as the second one (Fig. 11a). So far, only few atoms
on surfaces have been probed in this way, revealing interesting Kondo effects [87] and
single-atom magnetic anisotropies [88]. The next scientific step is to pass from atoms to
molecules in order to observe richer physics and to modify the properties of the magnetic

Molecular spintronics

13

objects. Although isolated SMMs on gold have been obtained [19, 20, 21, 22], the rather
drastic experimental requirements, i.e. very low temperatures and high magnetic fields,
have not yet been achieved. The first theoretical work predicted that quantum tunneling of
the magnetization is detectable via the electric current flowing through the molecule [89],
allowing therefore the readout of the quantum dynamics of a single molecule.
Another possibility concerns break-junction devices [90], which integrate a gate electrode. Such a three-terminal transport device, called a molecular spin-transistor, is a single
electron transistor with nonmagnetic electrodes and a single magnetic molecule as the
island. The current passes through the magnetic molecule via the source and drain electrodes, and the electronic transport properties are tuned via a gate voltage Vg (Fig. 11b).
Similarly to molecular electronics, weak- and strong-coupling regimes can be distinguished,
depending on the coupling between molecule and electrodes.
In the weak-coupling limit charging effects dominate the transport. Transport takes
place when a molecular orbital is in resonance with the Fermi energy of the leads and
electrons can then tunnel through the energy barrier into the molecular level and out into
the drain electrode. The resonance condition is obtained by shifting the energy levels with
Vg and the measurements show Coulomb-blockade diamonds [91].
The experimental realization of this scheme has been achieved using Mn12 with thiolcontaining ligands (Fig. 11b), which bind the SMM to the gold electrodes with strong and
reliable covalent bonds [16]. An alternative route is to use short but weak-binding ligands [17]: in both cases, the peripheral groups act as tunnel barriers and help conserving
the magnetic properties of the SMM in the junction. As the electron transfer involves the
charging of the molecule, we must consider, in addition to the neutral state, the magnetic
properties of the negatively- and positively-charged species. This introduces an important
difference with respect to the homologous measurements on diamagnetic molecules, where
the assumption is often made that charging of the molecule does not significantly alter the
internal degrees of freedom [92]. Because crystals of the charged species can be obtained,
SMMs permit direct comparison between spectroscopic transport measurements and more
traditional characterization methods. In particular, magnetization measurements, electron
paramagnetic resonance, and neutron spectroscopy can provide energy level spacings and
anisotropy parameters. In the case of Mn12 , positively charged clusters possess a lower
anisotropy barrier [93]. As revealed by the first Coulomb-blockade measurements, the
presence of these states is fundamental to explain transport through the clusters [16, 17].
Negative differential conductance was found that might be due to the magnetic characteristics of SMMs.
Studies in magnetic field showed a first evidence of the spin transistor properties [17].
Degeneracy at zero field and nonlinear behavior of the excitations as a function of field are
typical of tunneling via a magnetic molecule. In these first studies, the lack of a hysteretic
response can be due, besides environmental effects [23], to the alternation of the molecules
during the grafting procedure, to the population of excited states with lower energy barriers,
or might also be induced by the source-drain voltage scan performed at each field value.
Theoretical investigations in the weak-coupling regime predict many interesting effects. For example, a direct link between shot noise measurements and the detailed microscopic magnetic structure of SMMs has been proposed [94], allowing the connection
of structural and magnetic parameters to the transport features and therefore a characterization of SMMs using transport measurements. This opens the way to rational design of
SMMs for spintronics and to test the physical properties of related compounds. The first
step in this direction has already been made by comparing the expected response of chem-

14

Wolfgang Wernsdorfer

ically related SMMs [95]. Note that this direct link cannot be established for nanoparticles
or quantum-dots (QDs) because they do not posses a unique chemical structure.
A complete theoretical analysis as a function of the angle between the easy axis of
magnetization and the magnetic field showed that the response persists whatever the orientation of the SMM in the junction and that even films of SMMs should retain many salient
properties of single-molecule devices [96, 97].
For strong electronic coupling between the molecule and the leads, higher-order tunnel
processes become important, leading to the Kondo effect [98, 99, 100, 101]. This regime
has been attained using paramagnetic molecules containing one [85] or two magnetic centers [86], but remains elusive for SMMs.
The first mononuclear magnetic molecule investigated (Fig. 11c) is a Co2+ ion bound
by two terpyridine ligands, TerPy, attached to the electrodes with chemical groups of variable length [85]. The system with the longer alkyl spacer, due to a lower transparency of
the barrier, displays Coulomb blockade diamonds, which are characteristic for the weak
coupling regime, but no Kondo peak. Experiments conducted as a function of magnetic
field reveal the presence of excited states connected to spin excitations, in agreement with
the effective S = 1/2 state usually attributed to Co2+ ions at low temperatures but a Land
factor g = 2.1 is found. This is unexpected for Co2+ ions, characterized by high spin-orbit
coupling and magnetic anisotropy, and this point needs further investigation. The same
complex with the thiol directly connected to the TerPy ligand (Fig. 11d) shows strong
coupling to the electrodes, with exceptionally high Kondo temperatures around 25 K [85].
Additional physical effects of considerable interest were obtained using a simple molecule containing two magnetic centers [86]. This molecule, the divanadium molecule (Fig. 11e),
was again directly grafted to the electrodes, so as to have the highest possible transparency [86]. The molecule can be tuned with the gate voltage Vg into two differently
charged states. The neutral state, due to antiferromagnetic coupling between the two magnetic centers, has S = 0, while the positively charged state has S = 1/2. Kondo features
are found, as expected [98, 99, 100, 101], only for the state in which the molecule has
a nonzero spin moment. This nicely demonstrates that magnetic molecules with multiple centers and antiferromagnetic interactions permit to switch the Kondo effect on and
off, depending on their charge state. The Kondo temperature is again exceptionally high,
exceeding 30 K, and its characterization as a function of Vg indicates that not only spin
but also orbital degrees of freedom play an important role on the Kondo resonance of single molecules. Molecular magnets, in which spin-orbit interaction can be tuned without
altering the structure [25], are appealing to investigate further this physics.
The Kondo temperatures observed in the two cases [85, 86] are much higher than those
obtained for QDs and carbon nanotubes [98, 99, 100, 101], and are extremely encouraging.
The study of the superparamagnetic transition of SMMs while in the Kondo regime thus
seems achievable, possibly leading to an interesting interplay of the two effects. In order
to observe the Kondo regime one might start with small SMMs [25, 102], with core states
more affected by the proximity of the leads and use short and strongly bridging ligands to
connect SMMs to the electrodes [19, 85].
Theoretical investigations have explored the rich physics of this regime [94, 103, 104],
revealing that the Kondo effect should even be visible in SMMs with S > 1/2 [94]. This
is in contrast to expectations for a system with an anisotropy barrier, where the blocked
spin should hinder cotunneling processes. However in SMMs, the presence of a transverse
anisotropy induces a Kondo resonance peak [94]. The observation of this new physical
phenomena should be possible because of the tunability of SMMs, allowing a rational

Molecular spintronics

15

Figure 12
Spin-valves based on molecular magnets. Yellow arrows represent the magnetization.
a) Parallel configuration of the magnetic source electrode (copper color) and molecular magnetization, with diamagnetic drain electrode (golden color). Spin-up majority carriers (thick green arrow)
are not affected by the molecular magnetization, while the spin-down minority carriers (thin blue arrow) are partially reflected back. b) Anti-parallel configuration: majority spin-up electrons are only
partially transmitted by the differently polarized molecule, while the minority spin-down electrons
pass unaffected. Assuming that the spin-up contribution to the current is larger in the magnetic contact, this configuration has higher resistance than that of the previous case. c) Theoretical schematic
of a spin-valve configuration with nonmagnetic metal electrodes [8] and d) proposed molecular magnet between gold electrodes: a conjugated molecule bridges the cobaltocene (red) and ferrocene
(blue) moieties [106].

choice of the physical parameters governing the tunneling process: low symmetry transverse terms are particularly useful, because selection rules apply for high symmetry terms.
The first theoretical predictions argued that the Kondo effect should be present only for
half-integer spin molecules. However the particular quantum properties of SMMs allow
for the Kondo effect even for integer spins. In addition, the presence of the so-called
Berry-phase interference [30, 31, 69], a geometrical quantum phase effect, can produce
not only one Kondo resonance peak, but a series of peaks as a function of applied magnetic
field [105]. These predictions demonstrate how the molecular nature of SMMs and the
quantum effects they exhibit differentiate them from inorganic QDs and nanoparticles and
should permit the observation of otherwise prohibited phenomena.
4.2 Molecular spin-valve
A molecular spin-valve (SV) [8] is similar to a spin transistor but contains at least
two magnetic elements (Fig. 12a-b). SVs change their electrical resistance for different
mutual alignments of the magnetizations of the electrodes and of the molecule, analogous to a polarizer-analyzer setup. Non-molecular devices are already used in hard disc
drives, owing to the giant- and tunnel-magnetoresistance effects. As good efficiency has
already been demonstrated for organic materials [6], molecular SVs are actively sought
after [107, 108]. As only few examples of molecular SVs exist [109, 110], the fundamental physics behind these devices remains largely unexplored and will likely be the focus
of considerable attention in the near future. The simplest SV consists of a diamagnetic

16

Wolfgang Wernsdorfer

Figure 13
Molecular double-dot devices. Magnetic molecules proposed for grafting on suspended carbon nanotubes connected to Pd electrodes (form left to right): a C60 fullerene including a
rare-earth atom, the Mn12 SMM and the rare-earth-based double-decker [Tb(phtalocyanine)2 ] SMM.
The gate voltage of the double-dot device is obtained by a doped Si substrate covered by a SiO2 insulating layer.

molecule in between two magnetic leads, which can be metallic or semiconducting. The
first experiments sandwitched a C60 fullerene between Ni electrodes, showing a very large
negative magnetoresistance effect [109]. Another interesting possibility is to use carbon
nanotubes connected with magnetic halfmetallic electrodes transforming spin information
into large electrical signals [106].
A SMM-based SV can have one or two magnetic electrodes (Fig. 12a-b), or the molecule
can possess two magnetic centers in between two non-magnetic leads (Fig. 12c-d), in a
scheme reminiscent of early theoretical models of SVs [8]. Molecules with two magnetic centers connected by a molecular spacer are well-known in molecular magnetism
and a double metallocene junction has been theoretically studied [106]. This seems a good
choice, as the metallocenes leave the d-electrons of the metals largely unperturbed.
Theory indicates that, when using SMMs, the contemporary presence, at high bias, of
large currents and slow relaxation will individuate a physically interesting regime [111,
112]. Only spins parallel to the molecular magnetization can flow through the SMM and
the current will display, for a time equivalent to the relaxation time, a very high spin polarization. For large currents this process can lead to a selective drain of spins with one
orientation from the source electrode, thus transferring a large amount of magnetic moment from one lead to the other. This phenomenon, due to a sole SMM, has been named
giant spin amplification [111] and offers a convenient way to read the magnetic state of the
molecule. The switching of the device seems more complicated, at first sight, involving
a two-step process that includes the application of a magnetic field and the variation of
the bias voltage. However, it has recently been suggested that the spin-polarized current
itself can be sufficient to switch the magnetization of a SMM [113]. The switching can be
detected in the current as a step if both leads are magnetic and have parallel magnetization
or as a sharp peak for the anti-parallel configuration.
4.3 Molecular multi-dot devices
A double-dot devices (Figure 13) is one possible route for molecular spintronics [1].
It is a three terminal device, where the current passes through a non-magnetic quan-

Molecular spintronics

17

tum conductor (quantum wire, nanotube, molecule, or quantum dot (QD)). The magnetic
molecule is only weakly coupled to the non-magnetic conductor but its spin can influence
the transport properties, permitting readout of the spin state with minimal back-action.
Several mechanisms can be exploited to couple the two systems. One appealing way is
to use a carbon nanotube as a detector of the magnetic flux variation, possibly using the
nanoSQUID [2]. Other possibilities involve the indirect detection of the spin state through
electrometry. Indeed, a non-magnetic quantum conductor at low temperatures behaves as a
QD for which charging processes become quantized, giving rise to Coulomb blockade and
Kondo effect depending on the coupling to the leads. Any slight change in the electrostatic
environment (controlled by the gate) can induce a shift of the Coulomb diamonds of the
device, leading to a conductivity variation of the QD at constant gate voltage. QDs are
therefore accurate electrometers. When the QD is coupled, even weakly, with a magnetic
object, due to the Zeeman energy the spin flip at non-zero field induces a change of the
electrostatic environment of the QD. This effect, called magneto-Coulomb effect, enables
therefore to detect the magnetization reversal of the molecule.
Another route is weak exchange or dipole coupling between the magnetic molecule
and the QD. It is interesting to probe these effects as a function of the number of trapped
electrons because odd or even number of electrons should lead to different couplings. The
main advantage of these schemes is that the coupling to the leads and the injected current
does not alter the magnetic properties of the molecule. Because coupling is small, these
devices might allow a non-destructive readout of the spin states.

5 Conclusion
In conclusion, molecular nanomagnets offer a unique opportunity to explore the quantum dynamics of a large but finite spin. We focused our discussion on the Fe8 molecular
nanomagnet because it is the first system where studies in the pure quantum regime were
possible. In the coming years, chemistry is going to play a major role through the synthesis of novel larger spin clusters with strong anisotropy [9]. The unique properties of
SMMs will soon lead to design the molecules for specific transport characteristics using
the flexibility of supramolecular chemistry. Important investigations concern the studies
of the quantum character of molecular clusters for applications like quantum computers.
The first implementation of Grovers algorithm with molecular nanomagnets has been proposed [13]. Antiferromagnetic systems have attracted an increasing interest. In this case
the quantum hardware is thought of as a collection of coupled molecules, each corresponding to a different qubit [14, 15, 114, 115]. In order to explore these possibilities, new
and very precise setups are currently built and new methods and strategies are developed.
The field of molecular nanomagnets evolves towards molecular electronics and spintronics,
which are both rapidly emerging fields of nanoelectronics with a strong potential impact
for the realization of new functions and devices helpful for information storage as well as
quantum information. New projects aim at the merging of the two fields by the realization
of molecular junctions that involve a molecular nanomagnet. In order to tackle the challenge of controlled connection at the single molecule level, molecular self assembly on
nanojunctions obtained by the technique of electromigration was used [3, 16, 17]. Futhermore, a new nano-SQUID with carbon nanotube Josephson junctions was developed [2],
which should be sensitive enough to study individual magnetic molecules that are attached
to the carbon nanotube. Such techniques will lead to enormous progress in the understand-

18

Wolfgang Wernsdorfer

ing of the electronic and magnetic properties of isolated molecular systems and they will
reveal intriguing new physics [1].
The author is indebted to F. Balestro, N. Bendiab, L. Bogani, E. Bonet, J.-P. Cleuziou,
E. Eyraud, D. Lepoittevin, L. Marty, C. Thirion. This work is partially financed by STEP
MolSpinQIP, ERC-Advanced Grant MolNanoSpin, ANR Pnano MolNanoSpin.

References
1

Bogani, L. and Wernsdorfer, W. (2008) Molecular spintronics using singlemolecule magnets, Nat. Mat. 7(3), 179186.

Cleuziou, J.-P., Wernsdorfer, W., Bouchiat, V., Ondarcuhu, T., and Monthioux, M.
(2006) Carbon nanotube superconducting quantum interference device, Nature
Nanotech. 1, 53.

Roch, N., Florens, S., Bouchiat, V., Wernsdorfer, W., and Balestro, F. (2008)
Quantum phase transition in a single-molecule quantum dot, Nature 453, 633
(2008).

Wolf, S. A., Awshalom, D. D., Buhrman, R. A., Daughton, J. M., von Molnr, S.,
Roukes, M. L., Chtchelkanova, A. Y., and Treger, D. M. (2001) Spintronics: a
spin-based electronics vision for the future, Science 294, 449 (2001).

Awschalom, D. D. and Flatt, M. M. (2007) Challenges for semiconductor spintronics, Nature Phys. 3, 153159.

Xiong, Z. H., Wu, D., Valy Vardeny, Z., and Shi, J. (2004) Giant magnetoresistance in organic spin-valves, Nature 427, 821824 (2004).

Tao, N. J. (2006) Electron transport in molecular junctions, Nature Nanotech. 1,


173181.

Sanvito, S. and Rocha, A. R. (2006) Molecular-spintronics: the art of driving spin


through molecules, J. Comput. Theor. Nanosci. 3, 624642.

Milios, C., Vinslava, A., Wernsdorfer, W., Moggach, S., Parsons, S., Perlepes, S.,
Christou, G., and Brechin, E. (2007) A record anisotropy barrier for a singlemolecule magnet, J. Am. Chem. Soc. 129, 2754.

10

Ardavan, A., Rival, O., Morton, J. J., Blundell, S. J., Tyryshkin, A. M., Timco,
G. A., and E., W. R. (2007) Will spin-relaxation times in molecular magnets
permit quantum information processing?, Phys. Rev. Lett. 98, 057201.

11

Carretta, S., Santini, P., Amoretti, G., Guidi, T., Copley, J. R., Qiu, Y., Caciuffo,
R., Timco, G. A., and E., W. R. (2007) Quantum oscillations of the total spin
in a heterometallic antiferromagnetic ring: Evidence from neutron spectroscopy,
Phys. Rev. Lett. 98, 167401.

12

Bertaina, S., Gambarelli, S., Mitra, T., Tsukerblat, B., Muller, A., and Barbara, B.
(2008) Quantum oscillations in a molecular magnet, Nature 453, 203.

Molecular spintronics

19

13

Leuenberger, M. N. and Loss, D. (2001) Quantum computing with molecular


magnets, Nature 410, 789.

14

Troiani, F., Affronte, M., Carretta, S., Santini, P., and Amoretti, G. (2005) Proposal
for quantum gates in permanently coupled antiferromagnetic spin rings without
need of local fields, Phys. Rev. Lett. 94, 190501.

15

Troiani, F., Ghirri, A., Affronte, M., Carretta, S., Santini, P., Amoretti, G., Piligkos,
S., Timco, G., and Winpenny, R. (2005) Molecular engineering of antiferromagnetic rings for quantum computation, Phys. Rev. Lett. 94, 207208.

16

Heersche, H. B., de Groot, Z., Folk, J. A., van der Zant, H. S. J., Romeike, C.,
Wegewijs, M. R., Zobbi, L., Barreca, D., Tondello, E., and Cornia, A. (2006)
Electron transport through single mn12 molecular magnets, Phys. Rev. Lett. 96,
206801.

17

Jo, M.-H., Grose, J., Baheti, K., Deshmukh, M., Sokol, J., Rumberger, E., Hendrickson, D., Long, J., Park, H., and Ralph, D. (2006) Signatures of molecular
magnetism in single-molecule transport spectroscopy, Nano Lett. 6, 2014.

18

Christou, G., Gatteschi, D., Hendrickson, D., and Sessoli, R. (2000) Singlemolecule magnets, MRS Bulletin 25, 66.

19

Cornia, A., Fabretti, A. C., Zobbi, L., Caneschi, A., Gatteschi, D., Mannini, M.,
and Sessoli, R. (2006) Preparation of novel materials using SMMs, Struct. Bond.
122, 133.

20

Fleury, B., Catala, L., Huc, V., David, C., Zhong, W., Jegou, P., Baraton, L., Albouy,
S. P. P., and Mallah, T. (2005) A new approach to grafting a monolayer of oriented
mn12 nanomagnets on silicon, Chem. Commun. , 2020.

21

Naitabdi, A., Bucher, J., Gerbier, P., Rabu, P., and Drillon, M. (2005) Selfassembly and magnetism of mn12 nanomagnets on native and functionalized gold
surfaces, Adv. Mater. 17, 1612.

22

Coronado, E., Forment-Aliaga, A., Gaita-Arino, A., Gimenez-Saiz, C., Romero,


F., and Wernsdorfer, W. (2004) Polycationic Mn12 single-molecule magnets as
electron reservoirs with s > 10 ground states, Angew. Chem. Int. Ed. Engl. 43,
6152.

23

Bogani, L., Cavigli, L., Gurioli, M., Novak, R., Mannini, M., Caneschi, A., Pineider, F., Sessoli, R., Clemente-Leon, Coronado, E., Cornia, A., and Gatteschi,
D. (2007) Magneto-optical investigations of nanostructured materials based on
single-molecule magnets monitor strong environmental effects, Adv. Mater. 19,
3906.

24

Ishikawa, N., Sugita, M., and Wernsdorfer, W. (2005) Nuclear spin driven
quantum tunneling of magnetization in a new lanthanide single-molecule magnet:
Bis(phthalocyaninato)holmium anion, J. Am. Chem. Soc. 127, 3650.

25

Ishikawa, N., Sugita, M., and Wernsdorfer, W. (2005) Quantum tunneling of magnetization in lanthanide single-molecule magnets: Bis(phthalocyaninato)terbium
and bis(phthalocyaninato)dysprosium anions, Angew. Chem. Int. Ed. 44, 2.

20

Wolfgang Wernsdorfer

26

Friedman, J. R., Sarachik, M. P., Tejada, J., and Ziolo, R. (1996) Macroscopic
measurement of resonant magnetization tunneling in high-spin molecules, Phys.
Rev. Lett. 76, 3830.

27

Thomas, L., Lionti, F., Ballou, R., Gatteschi, D., Sessoli, R., and Barbara, B. (1996)
Macroscopic quantum tunneling of magnetization in a single crystal of nanomagnets, Nature (London) 383, 145147.

28

Sangregorio, C., Ohm, T., Paulsen, C., Sessoli, R., and Gatteschi, D. (1997) Quantum tunneling of the magnetization in an iron cluster nanomagnet, Phys. Rev. Lett.
78, 4645.

29

Wernsdorfer, W., Murugesu, M., and Christou, G. (2006) Resonant tunneling


in truly axial symmetry mn12 single-molecule magnets: Sharp crossover between
thermally assisted and pure quantum tunneling, Phys. Rev. Lett. 96, 057208.

30

W.Wernsdorfer and Sessoli, R. (1999) Quantum phase interference and parity


effects in magnetic molecular clusters, Science 284, 133.

31

Wernsdorfer, W., Chakov, N. E., and Christou, G. (2005) Quantum phase interference and spin-parity in mn12 single-molecule magnets, Phys. Rev. Lett. 95,
037203.

32

Lis, T. (1980) Preparation, structure, and magnetic properties of a dodecanuclear


mixed-valence manganese carboxylate, Acta Cryst. B 36, 2042.

33

Wieghardt, K., Pohl, K., Jibril, I., and Huttner, G. (1984) Hydrolysis products of
the monomeric amine complex (C6 H15 N3 )FeCl3 : The structure of the octameric
Iron(III) cation of [Fe8 ...], Angew. Chem. Int. Ed. Engl. 23, 7778.

34

Caneschi, A., Gatteschi, D., Laugier, J., Rey, P., Sessoli, R., and Zanchini, C.
(1991) Alternating current susceptibility, high field magnetization, and millimeter
band EPR evidence for a ground S=10 state in [Mn12 ...], J. Am. Chem. Soc. 113,
58735874.

35

Sessoli, R., Tsai, H.-L., Schake, A. R., Wang, S., Vincent, J. B., Folting, K., Gatteschi, D., Christou, G., and Hendrickson, D. N. (1993) High-spin molecules:
[Mn12 O12 (O2 CR)16 (H2 O)4 ], J. Am. Chem. Soc. 115, 18041816.

36

Sessoli, R., Gatteschi, D., Caneschi, A., and Novak, M. A. Magnetic bistability in
a metal-ion cluster, Nature 365, 141143.

37

Barra, A.-L., Debrunner, P., Gatteschi, D., Schulz, C. E., and Sessoli, R. (1996)
Superparamagnetic-like behavior in an octanuclear iron cluster, EuroPhys. Lett.
35, 133.

38

Novak, M. and Sessoli, R. Ac susceptibility relaxation studies on a manganese organic cluster compound: Mn12 -ac, (1995) In Quantum Tunneling
of Magnetization-QTM94, Gunther, L. and Barbara, B., editors, volume 301 of
NATO ASI Series E: Applied Sciences, 171188. Kluwer Academic Publishers,
London.

Molecular spintronics

21

39

Paulsen, C. and Park, J.-G. (1995) In Quantum Tunneling of MagnetizationQTM94, Gunther, L. and Barbara, B., editors, volume 301 of NATO ASI Series
E: Applied Sciences, 189205. Kluwer Academic Publishers, London.

40

Abragam, A. and Bleaney, B. (1970) Electron paramagnetic resonance of transition ions, Clarendon Press, Oxford.

41

Villain, J., Hartmann-Boutron, F., Sessoli, R., and Rettori, A. (1994) Magnetic
relaxation in big magnetic molecules, EuroPhys. Lett. 27, 159.

42

Politi, P., Rettori, A., Hartmann-Boutron, F., and Villain, J. (1995) Tunneling in
mesoscopic magnetic molecules, Phys. Rev. Lett. 75, 537.

43

Hartmann-Boutron, F., Politi, P., and Villain, J. (1996) Tunneling and magnetic
relaxation in mesoscopic molecules, Int. J. Mod. Phys. 10, 2577.

44

Villain, J., Wurger, A., Fort, A., and Rettori, A. (1997) Tunnel effect in magnetic
systems: from the microscopic description to the master equation, J. de Phys. I 7,
1583.

45

Garanin, D. and Chudnovsky, E. (1997) Thermally activated resonant magnetization tunneling in molecular magnets: Mn12ac and others, Phys. Rev. B 56,
11102.

46

Fort, A., Rettori, A., Villain, J., Gatteschi, D., and Sessoli, R. (1998) Mixed
quantum-thermal relaxation in mn12 acetate molecules, Phys. Rev. Lett. 80, 612.

47

Leuenberger, M. N. and Loss, D. (2000) Spin tunneling and phonon-assisted


relaxation in mn12 -acetate, Phys. Rev. B 61, 1286.

48

Ohm, T., Sangregorio, C., and Paulsen, C. (1998) Local field dynamics in a
resonant quantum tunneling system of magnetic molecules, Euro. Phys. J. B 6,
195.

49

Perenboom, J., Brooks, J., Hill, S., Hathaway, T., and Dalal, N. (1998) Relaxation
of the magnetization of Mn12 acetate, Phys. Rev. B 58, 330338.

50

Kent, A., Zhong, Y., Bokacheva, L., Ruiz, D., Hendrickson, D., and Sarachik,
M. (2000) Low-temperature magnetic hysteresis in Mn12 acetate single crystals,
EuroPhys. Lett. 49, 521527.

51

Caneschi, A., Gatteschi, D., Sangregorio, C., Sessoli, R., Sorace, L., Cornia, A.,
Novak, M. A., Paulsen, C., and Wernsdorfer, W. (1999) The molecular approach
to nanoscale magnetism, J. Magn. Magn. Mat. 200, 182.

52

Aubin, S. M. J., Dilley, N. R., Wemple, M. B., Christou, G., and Hendrickson,
D. N. (1998) Resonant magnetization tunnelling in the half-integer-spin singlemolecule magnet [PPh4 ][Mn12 O12 (O2 CEt)16 (H2 O)4 ], J. Am. Chem. Soc. 120,
839.

53

Price, D. J., Lionti, F., Ballou, R., Wood, P., and Powell, A. K. (1999) Large metal
clusters and lattices with analogues to biology, Phil. Trans. R. Soc. Lond. A 357,
3099.

22

Wolfgang Wernsdorfer

54

Yoo, J., Brechin, E. K., Yamaguchi, A., Nakano, M., Huffman, J. C., Maniero,
A., Brunel, L.-C., Awaga, K., Ishimoto, H., Christou, G., and Hendrickson, D. N.
(2000) Single-molecule magnets; a new class of tetranuclear magnganese magnets, Inorg. Chem. 39, 3615.

55

Tasiopoulos, A., Vinslava, A., Wernsdorfer, W., Abboud, K., and Christou, G.
(2004) Giant single-molecule magnets: A mn84 torus and its supramolecular nanotubes, Angew. Chem. Int. Ed. Engl. 43, 2117.

56

Pontillon, Y., Caneschi, A., Gatteschi, D., Sessoli, R., Ressouche, E., Schweizer,
J., and Leli`evre-Berna, E. (1999) Magnetization desnsity in an irob(iii) magnetic
cluster. a polarized neutron investigation, J. Am. Chem. Soc. 121, 5342.

57

Landau, L. (1932) On the theory of transfer of energy at collisions ii, Phys. Z.


Sowjetunion 2, 46.

58

Zener, C. (1932) Non-adiabatic crossing of energy levels, Proc. R. Soc. London,


Ser. A 137, 696.

59

St
uckelberg, E. (1932) Theorie der unelastischen stoesse zwischen atomen, Helv.
Phys. Acta 5, 369.

60

Miyashita, S. (1995) Dynamics of the magnetization with an inversion of the


magnetic field, J. Phys. Soc. Jpn. 64, 3207.

61

Miyashita, S. (1996) Observation of the energy gap due to the quantum tunneling
making use of the landau-zener mechanism [uniaxial magnets], J. Phys. Soc. Jpn.
65, 2734.

62

G.Rose and Stamp, P. (1998) Short time ac response of a system of nanomagnets,


Low Temp. Phys. 113, 1153.

63

Thorwart, M., Grifoni, M., and Hanggi, P. (2000) Strong coupling theory for
driven tunneling and vibrational relaxation, Phys. Rev. Lett. 85, 860.

64

Leuenberger, M. N. and Loss, D. (2000) Incoherent zener tunneling and its application to molecular magnets, Phys. Rev. B 61, 12200.

65

Wernsdorfer, W., Ohm, T., Sangregorio, C., Sessoli, R., Mailly, D., and Paulsen,
C. (1999) Observation of the distribution of molecular spin states by resonant
quantum tunneling of the magnetization, Phys. Rev. Lett. 82, 3903.

66

Loss, D., DiVincenzo, D. P., and Grinstein, G. (1992) Suppression of tunneling


by interference in half-integer-spin particles, Phys. Rev. Lett. 69, 3232.

67

von Delft, J. and Henley, C. L. (1992) Destructive quantum interference in spin


tunneling problems Phys. Rev. Lett. 69, 3236 (1992).

68

Feynman, R. P., Leighton, R. B., and Sand, M. (1970) The Feynman lectures on
physics, volume 3 Addison-Wesley Publishing Company, London.

69

Garg, A. (1993) Topologically quenched tunnel splitting in spin systems without


kramers degeneracy, EuroPhys. Lett. 22, 205.

Molecular spintronics

23

70

Garg, A. (1999) Oscillatory tunnel splittings in spin systems: A discrete wentzelkramers-brillouin approach, Phys. Rev. Lett. 83, 4385.

71

Villain, J. and Fort, A. (2000) Magnetic tunneling told to ignorants by two ignorants, Euro. Phys. J. B 17, 69.

72

Kececioglu, E. and Garg, A. (2001) Diabolical points in magnetic molecules: An


exactly solvable model, Phys. Rev. B 63, 064422.

73

Barnes, S. (1999) Manifestation of intermediate spin for fe8 , cond-mat/9907257.

74

Liang, J.-Q., Mueller-Kirsten, H., Park, D., and Pu, F.-C. (2000) Phase interference for quantum tunneling in spin systems, Phys. Rev. B 61, 8856.

75

Yoo, S.-K. and Lee, S.-Y. (2000) Geometrical phase effects in biaxial nanomagentic particles, Phys. Rev. B 62, 3014.

76

L
u, R., Hu, H., Zhu, J.-L., Wang, X.-B., Chang, L., and Gu, B.-L. (2000) Topological phase interference induced by a magnetic field along hard anisotropy axis
in nanospin systems with different crystal symmetries, Phys. Rev. B 61, 14581.

77

L
u, R., Kou, S.-P., Zhu, J.-L., Chang, L., and Gu, B.-L. (2000) Magnetization
quantum tunneling at excited levels for a biaxial spin system in an arbitrarily directed magnetic field, Phys. Rev. B 62, 3346.

78

Caciuffo, R., Amoretti, G., Murani, A., Sessoli, R., Caneschi, A., and Gatteschi, D.
(1998) Neutron spectroscopy for the magnetic anisotropy of molecular clusters,
Phys. Rev. Lett. 81, 4744.

79

Amoretti, G., Caciuffo, R., Combet, J., Murani, A., and Caneschi, A. (2000) Inelastic neutron scattering below 85 mev and zero-field splitting parameters in the
fe8 magnetic cluster, Phys. Rev. B 62, 3022.

80

Barra, A., Gatteschi, D., and Sessoli, R. (2000) High-frequency EPR spetra of
Fe8. a critical appraisal of the barrier for the reorientation of the magnetization in
single-molecule magnets, Chem. Eur. J. 6, 1608.

81

Enz, M. and Schilling, R. (1986) Magnetic field dependence of the tunnelling


splitting of quantum spins, J. Phys. C 19, L711.

82

Hemmen, J. L. V. and S
uto, S. (1986) Tunneling of quantum spins, EuroPhys.
Lett. 1, 481.

83

Kramers, H. A. (1930) Proc. Acad. Sci. Amsterdam 33, 959.

84

Wernsdorfer, W., Bhaduri, S., Boskovic, C., Christou, G., and Hendrickson, D.
(2002) Spin-parity dependent tunneling of magnetization in single-molecule magnets, Phys. Rev. B 65, 180403(R).

85

Park, J., Pasupathy, A. N., Goldsmith, J. I., Chang, C., Petta, Y. Y. J. R., Rinkoski,
M., Sethna, J. P., Abruna, H. D., McEuen, P. L., and Ralph, D. C. (2002) Coulomb
blockade and the kondo effect in single-atom transistors, Nature 417, 722 (2002).

24

Wolfgang Wernsdorfer

86

Liang, W., Bockrath, M., and Park, H. (2002) Shell filling and exchange coupling
in metallic single-walled carbon nanotubes, Phys. Rev. Lett. 88(12), 126801.

87

Wahl, P., Simon, P., Diekhoner, L., Stepanyuk, V., Bruno, P., Schneider, M., and
Kern, K. (2007) Exchange interaction between single magnetic adatoms, Phys.
Rev. Lett. 98, 056601.

88

Hirjibehedin, C. F., Lin, C. Y., Otte, A. F., Ternes, M., Lutz, C. P., Jones, B. A.,
and Heinrich, A. J. (2007) Large magnetic anisotropy of a single atomic spin
embedded in a surface molecular network, Science 317, 1199.

89

Kim, G. and Kim, T. (2004) Electronic transport in single molecule magnets on


metallic surfaces, Phys. Rev. Lett. 92, 137203.

90

Park, H., Kim, A. K. L., Alivisatos, A. P., Park, J., and McEuen, P. L. (1999)
Fabrication of metallic electrodes with nanometer separation by electromigration,
Appl. Phys. Lett. 75, 301.

91

Hanson, R., Kouwenhoven, L., Petta, J., Tarucha, S., and Vandersypen, L. (2007)
Spins in few-electron quantum dots, Rev. Mod. Phys. 79, 1217.

92

Kouwenhoven, L. P., Marcus, C. M., McEuen, P. L., Tarucha, S., and Wingreen, R.
M. W. N. (1997) Electron transport in quantum dots, In Mesoscopic Electron
Transport, Sohn, L. L., Kouwenhoven, L. P., and Schoen, G., editors, volume 345
of Series E, 105214. Kluwer, London.

93

Chakov, N. E., Soler, M., Wernsdorfer, W., Asbboud, K. A., and Christou, G.
(2005) Electron reduction in mn12 single-molecule magnets: Structural characterization, magnetic properties and 19 f nuclear magnetic resonance spectroscopy
of a family of mn12 clusters spanning three oxidation levels, Inorg. Chem. 44,
53045321.

94

Romeike, C., Wegewijs, M. R., and Schoeller, H. (2006) Spin quantum tunneling
in single molecule magnets: Fingerprints in transport spectroscopy of current and
noise, Phys. Rev. Lett. 96, 196805.

95

Romeike, C., Wegewijs, R., Park, M. R., Ruben, M., Wenzel, W., and Schoeller, H.
(2007) Charge-switchable molecular nanomagnet and spin-blockade tunneling,
Phys. Rev. B 75, 064404.

96

Elste, F. and Timm, C. (2007) Cotunneling and nonequilibrium magnetization in


magnetic molecular monolayers, Phys. Rev. B 75, 195341.

97

Timm, C. (2007) Tunneling through magnetic molecules with arbitrary angle


between easy axis and magnetic field, Phys. Rev. B 76, 014421.

98

Goldhaber-Gordon, D., Gores, J., Kastner, M. A., Shtrikman, H., Mahalu, D., and
Meirav, U. (1998) From the kondo regime to the mixed-valence regime in a
single-electron transistor, Phys. Rev. Lett. 81(23), 5225.

99

Cronenwett, S. M., Oosterkamp, T. H., and Kouwenhoven, L. P. (1998) A tunable


kondo effect in quantum dots, Science 281, 540.

Molecular spintronics

25

100

van der Wiel, W. G., Franceschi, S. D., Fujisawa, T., Elzerman, J. M., Tarucha, S.,
and Kouwenhoven, L. P. (2000) The kondo effect in the unitary limit, Science
289, 2105.

101

Nygard, J., Cobden, D. H., and Lindelof, P. E. (2000) Kondo physics in carbon
nanotubes, Nature 408, 342 (2000).

102

Aromi, G. and Brechin, E. K. (2006) Synthesis of 3d metallic single-molecule


magnets, Struct. Bond. 122, 167.

103

Romeike, C., Wegewijs, M. R., and Schoeller, H. (2006) Quantum-tunnelinginduced kondo effect in single molecular magnets, Phys. Rev. Lett. 96, 196601.

104

Romeike, C., Wegewijs, M. R., and Schoeller, H. (2006) Kondo-transport spectroscopy of single molecule magnets, Phys. Rev. Lett. 97, 206601.

105

Leuenberger, M. N. and Mucciolo, E. R. (2006) Berry-phase oscillations of the


kondo effect in single-molecule magnets, Phys. Rev. Lett. 97, 126601.

106

Liu, R., Ke, S.-H., Baranger, H. U., and Yang, W. (2006) Negative differential
resistance and hysteresis through an organometallic molecule from molecular-level
crossing, J. Am. Chem. Soc. 128, 6274.

107

Waldron, D., Haney, P., Larade, B., MacDonald, A., and Guo, H. (2006) Nonlinear
spin current and magnetoresistance of molecular tunnel junctions, Phys. Rev. Lett.
96, 166804.

108

Rocha, A. R., Garcia-Suarez, V. M., Bailey, S. W., Lambert, C. J., Ferrer, J., and
Sanvito, S. (2005) Towards molecular spintronics, Nat. Mat. 4, 335.

109

Pasupathy, A., Bialczak, R., Martinek, J., Grose, J., Donev, L., McEuen, P., and
Ralph, D. (2004) The kondo effect in the presence of ferromagnetism, Science
306, 86.

110

Hueso, L., Pruneda, J. M., Ferrari, V., Burnell, G., Valds-Herrera, J. P., Simons,
B. D., Littlewood, P. B., Artacho, E., Fert, A., and Mathur, N. D. (2007) Transformation of spin information into large electrical signals using carbon nanotubes,
Nature 445, 410.

111

Timm, C. and Elste, F. (2006) Spin amplification, reading, and writing in transport
through anisotropic magnetic molecules, Phys. Rev. B 73, 235304.

112

Elste, F. and Timm, C. (2006) Transport through anisotropic magnetic molecules


with partially ferromagnetic leads: Spin-charge conversion and negative differential
conductance, Phys. Rev. B 73, 235305.

113

Misiorny, M. and Barnas, J. (2007) Magnetic switching of a single molecular


magnet due to spin-polarized current, Phys. Rev. B 75, 134425.

114

Meier, F., Levy, J., and Loss, D. (2003) Magnetization transport and quantized
spin conductance, Phys. Rev. Lett. 90, 047901.

115

Meier, F., Levy, J., and Loss, D. (2003) Quantum computing with antiferromagnetic spin clusters, Phys. Rev. B 68, 134417.

Vous aimerez peut-être aussi