Вы находитесь на странице: 1из 13

PAPER www.rsc.

org/materials | Journal of Materials Chemistry

Ordered mesoporous silicas as host for the incorporation and aggregation of


octanuclear nickel(II) single-molecule magnets: a bottom-up approach to
new magnetic nanocomposite materials{{
Emilio Pardo,a Pedro Burguete,b Rafael Ruiz-Garcı́a,a Miguel Julve,a Daniel Beltrán,b Yves Journaux,c
Pedro Amorós*b and Francesc Lloret*a
Received 24th February 2006, Accepted 13th April 2006
First published as an Advance Article on the web 11th May 2006
DOI: 10.1039/b602838a

Silica-based mesoporous materials have been employed as the support host for a suitably designed
small octanuclear nickel(II) guest complex with a moderately anisotropic S = 4 ground spin state
(D = 20.23 cm21), which behaves as a single-molecule magnet at low temperature (TB = 3.0 K).
Both unimodal MCM-41 and bimodal UVM-7 porous silica provide appropriate template
conditions for the incorporation and aggregation of the Ni8 complex precursor into larger
complex aggregates, showing slow relaxation of the magnetization at higher blocking
temperatures than the crystalline material. By playing with the initial complex vs. silica
concentration, two series of samples with varying complex loading amounts have been obtained.
The degree of aggregation varies, largely depending on the silica used, being higher for the
bimodal UVM-7 silica series. The mesophasic and porous nature of the Ni8 adsorbed silica
samples has been verified from XRD and TEM images. N2 adsorption–desorption isotherms
show that incorporation initiates inside the small intra-particle mesopores while subsequent
aggregation occurs at the external particle surface (close to the mesopore entrances). Both DC and
AC magnetic susceptibility measurements have demonstrated the occurrence of such a unique
silica-mediated surface aggregation process of cationic Ni8 molecules into oligomeric [Ni8]x
aggregates of large spin values (S = 4x) and high blocking temperatures (TB = 4.5–10.5 K). The
existence of a wide distribution of aggregates with different conformation and association degree
(size distribution) and the presence of weak interactions between the aggregates leads to an exotic
spin glass magnetic behavior for this family of host–guest hybrid nanocomposite materials.

Introduction promising support family for guest molecules due to their


high surface area and large pore volume. Since its discovery
The morphosynthesis of novel materials from the multilevel in 1992,3 the technological importance of M41 silicas and
organization of molecules with interesting physical and related derivatives have sparked extensive research interest
chemical properties into suitable ordered host systems is a in materials science. The number of reports dealing with the
major current challenge in materials science and nanotechno- use of mesoporous silicas as inorganic hosts for the incorpora-
logy.1 This molecular ‘bottom-up’ approach appears as an tion of magnetic guest molecules is relatively scarce when
alternative to the classical ‘top-down’ one to creating nano- compared with other typical applications of M41s and related
scopic functional materials displaying redox, magnetic, optical, materials.4,5
sensor, or catalytic activities.2 Organic (polymers) and Among the magnetic guest molecules, transition metal
inorganic (ceramics, porous solids) materials have been utilized polynuclear coordination compounds have attracted much
as support host systems. Among porous inorganic materials, attention since the discovery, in 1993, of a dodecanuclear
ordered mesoporous silicas have received much attention as a mixed-valence oxomanganese(III,IV) complex that behaves
as a molecular nanomagnet, termed single-molecule magnet
a
Instituto de Ciencia Molecular (ICMOL), Universitat de València, (SMM).6 This compound of formula Mn12O12(O2CMe)16-
46100 Burjassot, Valencia, Spain. E-mail: francisco.lloret@uv.es; (H2O)4, referred to as Mn12, exhibits a slow relaxation of the
Fax: 34 96354 4322; Tel: 34 96354 4441
b
Institut de Ciència dels Materials (ICMUV), Universitat de València, magnetization below a blocking temperature (TB) of 4.0 K,
P. O. Box 2085, 46071 Valencia, Spain. E-mail: pedro.amoros@uv.es; similar to that of superparamagnetic nanoparticles of common
Fax: 34 96354 3633; Tel: 34 96354 3617 magnetic materials such as metals or metal oxides. The
c
Laboratoire de Chimie Inorganique et Matériaux Moléculaires, UMR
CNRS 7071, Université Pierre et Marie Curie, 75252 Paris, France
combination of a large spin ground state (S = 10) with a large
{ Electronic Supplementary Information (ESI) available: Fig. S1: and negative axial magnetic anisotropy (D = 20.5 cm21) in
Temperature dependence of xM9 and xM0 for 1A–I; Fig. S2: Arrhenius Mn12 results in an effective energy barrier (Ueff = 2DS2 =
and Vogel–Fulcher plots for 1A–1G and 1I. See http://dx.doi.org/ 50 cm21) for the magnetization reversal between the two
10.1039/b602838a/.
{ This article is part of a themed issue on Molecular Magnetic lowest mS = ¡S states which accounts for the observed
Materials. magnetic bistability.7

2702 | J. Mater. Chem., 2006, 16, 2702–2714 This journal is ß The Royal Society of Chemistry 2006
Addressing individual SMMs or their aggregates, as well as
increasing their blocking temperatures, are conditions to be
fulfilled for potential applications of such molecule-based
nanoscopic magnetic materials. The recent success in the
elaboration of oxomanganese SMMs of large dimensions by
Christou and co-workers is illustrated by the nanometer-sized
Mn30 and Mn84 complexes prepared from the self-aggregation
in solution of smaller-sized Mn12 SMMs under different
reaction conditions.8 Nevertheless, these giant SMMs
possess TB values of only 1.5 K due to the relatively small
spin ground state (S = 5–6) in spite of the moderately large
magnetic anisotropy (D # 20.5 cm21). In fact, these
rather serendipitous, homogeneous self-assembly methods are
severely limited by the lack of control over the structure and
topology of the final complex, properties that ultimately
determine both S and D values of SMMs.
By following a completely different heterogeneous self-
assembly approach, several groups have recently reported the
use of both inorganic and organic materials as support host
systems for the organization of SMMs.9–15 These include a
variety of surfaces and other confined media such as multi-
layered Langmuir–Blodgett films,9a mesoporous silicas,9b,9c,10
polycarbonate thin films,11a,11c silicon surfaces,11b,12b highly
oriented pyrolytic graphite (HOPG),11d gold films,9d,12a,13 self-
assembled monolayers (SAMs),14 and ethyl acrylate poly-
mers.15 There are not a large variety of polynuclear complex
types, and in practically all cases the studies are devoted to the
incorporation of Mn12 SMMs with differently functionalized
carboxylate ligands, which are bonded to the surface through
electrostatic interactions, p–p interactions, hydrogen bonds, or
covalent bonds. Although the possibility of addressing isolated
or small nanometer-sized aggregates of Mn12 molecules (from
one to a few hundred) has been demonstrated, increase of the
blocking temperature has not been achieved by either of the
ordered hosts used. In all the cases, the magnetic properties of
the incorporated neutral Mn12 molecules (whenever reported)
are practically identical to those of the isolated molecules in
the crystalline material.9a–c,10,11a,15
In this article, we present a simple synthetic strategy to
prepare new magnetic nanocomposite materials from the
heterogeneous chemical aggregation of preformed polynuclear
building blocks, other than those of the Mn12 family, that
behave themselves as SMMs through the use of ordered
mesoporous silica as host matrix. As the magnetic guest
molecule, we have selected the octanuclear nickel(II) oxamate
complex of formula {[Ni2(mpba)3][Ni(dpt)(H2O)]6}(ClO4)4?
12.5H2O 1, where mpba is the bridging ligand m-phenylene-
dioxamate and dpt is the terminal ligand dipropylenetriamine Scheme 1 (a) Side and top views of the molecular structure and the
(Scheme 1).16 This complex has been prepared through a step- spin topology of the cationic Ni8 cage complex 1. (b) Structural
by-step strategy (the so-called ‘complex-as-ligand’ approach) formula of the bridging mpba and terminal dpt ligands.
which allows the rational design of metal coordination cages
from self-assembled, exchange-coupled metallacyclic com- complex consists of two strong antiferromagnetically coupled,
plexes with aromatic oligooxamate ligands.17 Complex 1 oxamate-bridged, propeller-like Ni4 units (J = 226.6 cm21)
possesses a moderately anisotropic S = 4 ground spin state which are weakly ferromagnetically coupled through the meta-
(D = 20.23 cm21) and exhibits slow magnetic relaxation below substituted phenylenediamidate bridges (J9 = +3.1 cm21)
TB = 3.0 K in the crystalline state. The structure 1 is built from (Ni–Ni separations of ca. 6.8 and 5.5 Å, respectively). These
octanickel(II) cations with a dimer-of-tetramers structure of intra-molecular magnetic interactions lead to a peculiar spin
double-propeller topology, separated by perchlorate anions topology for this Ni8 molecule with a ‘spin down’ orientation
and water molecules of crystallization. Each cationic Ni8 cage of the two central high-spin NiII ions and a ‘spin up’

This journal is ß The Royal Society of Chemistry 2006 J. Mater. Chem., 2006, 16, 2702–2714 | 2703
orientation of the six peripheral high-spin NiII ions dissolved in 5 ml of a basic water solution (pH = 8.5) (adjusted
(Scheme 1a). Importantly, one coordination position of each by addition of NaOH), giving a pale blue solution. Then,
of the six peripheral octahedral metal ions, partially blocked 100 mg of MCM-41 mesoporous silica were added and
with the dipropylenetriamine ligand, is occupied by a labile the reaction mixture was stirred during 30 min at room
water molecule that can be readily exchanged by anchoring temperature to favor the incorporation of the complex into the
and/or bridging deprotonated silanol groups from the host silica. The resulting mesostructured light-blue powder was
matrix during the process of mesoporous incorporation. separated by filtration, washed with water and hexane, and air
Herein, we report on an extensive series of host–guest hybrid dried. Samples 1B and 1C were synthesized following the same
nanocomposite materials prepared from the self-organization procedure by starting from solutions containing 40 and 60 mg
of a self-assembled metallosupramolecular Ni8 complex pre- of complex 1, respectively.
cursor into higher nuclearity aggregates within different SiO2
porous hosts. These include two ordered mesoporous silica Ni8 adsorbed UVM-7 nanocomposites (1D–H). Samples 1D
supports, either unimodal or bimodal like MCM-41 and to 1H were synthesized following the aforementioned proce-
UVM-7, respectively, as well as a non-ordered porous xerogel dure by starting from solutions containing 100 mg of UVM-7
denoted UVM-11. UVM-7 silica consists of MCM-41 type mesoporous silica and 20, 40, 60, 80, and 100 mg of complex 1,
mesoporous nanosized particles (average diameter of ca. respectively.
15–20 nm) joined together in micrometric aggregates (from
0.5 to 2 mm). In addition to the high surface area typical of M41 Ni8 adsorbed UVM-11 nanocomposites (1I). Sample 1I was
silicas, this special topology provides a significant intraparticle synthesized following the aforementioned procedure by start-
mesopore length decrease (from mm to nm) when compared ing from a solution containing 100 mg of UVM-11 xerogel and
with MCM-41 solids, together with an important interparticle 60 mg of complex 1.
textural porosity in the border between meso and macropores Table 1 summarizes the initial complex 1 concentration
that confers on UVM-7 materials their hierarchical porous used for the preparation of samples 1A–I, along with selected
nature. This architecture minimizes possible pore blocking physical data for each series of samples compared with those
phenomena inside the intraparticle mesopores by guest species, of their corresponding host.
and enhances at the same time the accessibility to the mesopore
entrances through the large interparticle pore system. Taking Physical techniques
advantage of the different chemical and physical properties of
these silica materials as host matrix, we try to give further The chemical composition of samples 1A–I were determined
insights into the mechanism of such a unique mesoporous by elemental analysis (C, H, N) and electron probe micro-
silica-mediated aggregation process of cationic Ni8 molecules analysis (EPMA) (Philips SEM-515). The preservation of the
into oligomeric [Ni8]x aggregates of larger spin values and nanoscopic organization typical of ordered mesoporous silicas
higher blocking temperatures. We show that the basic chemical was confirmed by X-ray diffraction (XRD) (Seifert 3000TT h-h
integrity of the octanuclear complex precursor is retained, diffractometer using Cu–Ka radiation), transmission electron
while weak ferromagnetic exchange interactions are estab- microscopy (TEM) (JEOL JEM-1010 operating at 100 kV),
lished between the octanuclear units in the resultant aggregate, and N2 adsorption–desorption isotherms (ASAP2010 analy-
leading to rather exotic magnetic properties for this family of zer). Variable-temperature (1.8–300 K) direct current (DC)
Ni8 adsorbed nanocomposite materials. and alternating current (AC) magnetic susceptibility measure-
ments and variable-field (0–5 T) magnetization measurements
were carried out on powdered samples of 1A–I with a
Experimental Quantum Design SQUID magnetometer. The magnetic data
Materials were corrected for the diamagnetism of the silica content of the
samples and for the sample holder.
Complex 1 was prepared as reported previously.16 The silica-
based host materials, MCM-41, UVM-7 and UVM-11, were
obtained by using the so-called atrane route as described in
Results and discussion
detail elsewhere.18 This strategy consists in the use of silicon Synthesis and general physical characterization
atrane complexes (silatrane) as hydrolytic inorganic precursors
and surfactants as porogen species. MCM-41 and UVM-7 The influence of the nature of the porous host on the loading
silicas have been isolated through a surfactant-assisted pro- efficiency and the location preferences of the Ni8 molecules
cedure. Cetyltrimethylammonium bromide (CTABr) has been has been investigated for three different types of silica-based
used as template agent. Surfactant evolution was carried out materials: unimodal MCM-41 and bimodal UVM-7 meso-
under soft-chemical conditions (chemical exchange in EtOH– porous silicas, and nanoparticulated UVM-11 silica xerogels.
HCl mixture). UVM-11 xerogels have been prepared following In the former two cases, the adsorption of complex 1 within
a non-surfactant strategy also through the atrane route. mesoporous silicas is a very fast and highly efficient process
owing to the different net charge of the silica surface
Preparations possessing anionic silanolate groups and the cationic Ni8
species, which favors effective complex incorporation through
Ni8 adsorbed MCM-41 nanocomposites (1A–C). In a typical electrostatic interactions. The use of cationic surfactants
synthesis leading to sample 1A, 20 mg of complex 1 were together with the soft chemical exchange method employed

2704 | J. Mater. Chem., 2006, 16, 2702–2714 This journal is ß The Royal Society of Chemistry 2006
Table 1 Selected physical data for 1A–I compared with those of their corresponding host
[1]/weight Ni:Sia/molar BET Small pore Small pore Large pore Large pore
Sample ratio ratio area/m2 g21 sizeb/nm volumeb /cm3 g21 sizeb/nm volumeb/cm3 g21

MCM-41 874.2 2.80 0.63


1A 0.2 0.039 628.9 2.71 0.62
1B 0.4 0.065 147.9 0.12
1C 0.6 0.091 17.6 0.04
UVM-7 971.6 2.58 0.69 34.89 0.79
1D 0.2 0.032 166.4 0.15 30.93 0.78
1E 0.4 0.056 145.9 0.13 30.45 0.72
1F 0.6 0.080 92.5 0.08 31.57 0.40
1G 0.8 0.104 90.1 0.07 28.95 0.32
1H 1.0 0.120 90.4 0.06 29.35 0.29
UVM-11 145.9 30.45 0.72
1I 0.6 0.036 92.5 29.85 0.60
a b
Calculated from elemental analysis (C, H, N) and EPMA. Values estimated by applying the BJH model on the adsorption branch of
the isotherms.

for their evolution led to a high density of negatively charged molar ratio (sample 1C) (Table 1). At this same complex
silanol groups on the surface able to interact with complex 1. concentration, the UVM-11 silica shows an adsorption ability
We worked in a moderately basic medium (pH = 8.5) in order of only 0.036 Ni–Si molar ratio under saturation conditions
to maintain the cage complex structure in solution and to (sample 1I) (Table 1).
avoid excessive back-dissolution of the silica support while This behavior indicates a high affinity of UVM-7 and
maximizing the host/guest electrostatic interactions and, MCM-41 mesoporous silica surfaces for electrostatic cationic
consequently, the complex loading amount. Under these complex immobilization, which contrasts with the moderate to
conditions, the adsorption is complete in less than 30 min as low affinity of the UVM-11 silica surface. Indeed, all three
deduced from the disappearance of the pale blue colour of silica-based support materials present almost the same surface
complex 1 in solution, which acts as an indicator that informs density of silanol groups (ca. 5–6 Si–OH/nm2) estimated from
us about the total or partial complex incorporation. 29
Si NMR and porosimetry data. Thus, the differences in
The evolution of the complex 1 loading at varying initial complex 1 adsorption should have a topological or templating
concentration values for the different silica hosts is shown in origin which are related to the distinct active silica surfaces,
Fig. 1. UVM-7 silica presents the highest loading capability. In either intra- and/or inter-particle. The cylindrical shape of the
fact, the adsorption is complete (colorless mother solution Ni8 molecules with approximate height 6 diameter dimen-
after filtration) even at relatively high complex concentration sions of 1.5 6 2.0 nm is adequate to allow their hosting in both
in the range 20–80% weight (samples 1D–G). The highest pore systems. Complex diffusion through the small intra-
loading amount value of 0.120 Ni–Si molar ratio is achieved at particle UVM-7 and MCM-41 mesopores (ca. 2.5 nm) can be
a relatively high complex concentration of 100% weight under anticipated because their sizes are slightly greater than the
saturation conditions (sample 1H) (Table 1). The loading complex diameter. The fact that UVM-11 silica surface is
ability of MCM-41 silica is significantly lower and saturation rather inefficient for complex 1 adsorption, in spite of the
is achieved at an intermediate complex concentration of 60% similarity in the inter-particle pore morphology with UVM-7
weight, with a maximum loading amount value of 0.091 Ni–Si silica, underlines the importance of the intra-particle porosity
on the loading efficiency.
The chemical identity of the adsorbed cationic Ni8 species in
samples 1A–I was confirmed by elemental analysis (C, H, N)
and electron probe microanalysis (EPMA). The experimental
values of the C–N molar ratio in the range 2.6–2.9 for 1A–I
agree fairly well with the expected value for a 3:6 mpba:dpt
stoichiometry in complex 1 (C–N molar ratio = 2.75), thus
showing that the basic integrity of the adsorbed Ni8 cation is
retained. On the other hand, EPMA clearly shows the absence
of ClO42 anions in the final materials. Moreover, EPMA also
shows that all samples are chemically homogeneous (regular
distributions of Ni and Si atoms) at the micrometer level (spot
area ca. 1 mm) with constant and well-defined compositions. In
addition, the mesophasic and porous nature (mesopore
symmetry and occupancy) of the Ni8 adsorbed silica samples
has been verified from transmission electronic microscopy
images and X-ray powder diffraction, while N2 adsorption–
Fig. 1 Dependence of the Ni–Si content for 1A–I on the initial desorption isotherms provide valuable information on the
relative concentration of 1 (data from Table 1). location of the Ni8 molecules.

This journal is ß The Royal Society of Chemistry 2006 J. Mater. Chem., 2006, 16, 2702–2714 | 2705
Transmission electronic microscopy and X-ray powder dif-
fraction. The TEM images of samples 1A–I confirm that the
mesoscopic architecture of the three silica hosts is maintained
after complex 1 immobilization (Fig. 2). The existence of
aggregates constructed from nanoparticles, and the large inter-
particle pores defined among them, can be clearly observed
from TEM images of the Ni8 adsorbed nanocomposites
prepared from UVM-7 and UVM-11 hosts. The preservation
of the small mesopore system is appreciated in the enlarged
TEM images of Ni8 adsorbed UVM-7 and MCM-41
nanocomposites (insets of Fig. 2). Regardless of the silica host
type, TEM images allow us to discard the segregation of
complex 1 as independent crystals in the final materials with
confidence. Moreover, the absence of diffraction peaks at high
2h values in the XRD patterns of samples 1A–I confirms the Fig. 3 XRD patterns for samples 1D (a), 1E (b), 1G (c), and 1H (d).
absence of complex 1 as a secondary phase. The inset shows the XRD pattern of the parent UVM-7 pure silica.
The XRD patterns of samples 1D–H are characteristic of
UVM-7 type materials (Fig. 3). They display a strong in a more abrupt way (Fig. 4). Within this series, only sample
diffraction peak together with a small and broad signal in 1A maintains the low-angle signals typical of hexagonal
the low angle region. This pattern indicates the existence of a ordered mesoporous materials (inset of Fig. 4). Samples 1B
certain hexagonal order in the intra-particle mesopore (small and 1C show a complete phase cancellation instead. However,
pores) system. Thus, the order degree is maintained and the TEM images confirm the preservation of the hexagonal order
XRD peak positions remain practically unchanged after the in the mesopore array in these two samples also. Overall, XRD
incorporation of complex 1. The XRD patterns of the Ni8 data suggest a higher Ni8 inclusion inside the small mesopores
adsorbed UVM-7 nanocomposites differ only in the peak of MCM-41 silica when compared with UVM-7 host. This
intensities, which are significantly reduced when compared behavior could be attributed to the more regular and
with that of pure UVM-7 silica (inset of Fig. 3). The peak cylindrical shape of the MCM-41 mesopores, which should
intensity strongly drops when going from UVM-7 pure silica favor the diffusion of the Ni8 molecules along the pores.
to sample 1D and then it decreases in a smoother way for As expected, the Ni8 adsorbed UVM-11 nanocomposite
samples 1E to 1H. This intensity loss must be attributed to a (sample 1I) does not display XRD signals at all (not shown).
progressive phase cancellation phenomenon associated with
the introduction of scattering material into the mesopore. N2 adsorption–desorption. The N2 adsorption–desorption
Thus, XRD data confirm at least a partial pore filling by Ni8 isotherms of samples 1A–I are characteristic of their parent
molecules inside the small intra-particle mesopore system of silica-based porous materials (Figs. 5–7). The first adsorption
UVM-7. By contrast, the phase cancellation in the Ni8 step at intermediate P/P0 values for the Ni8 adsorbed MCM-41
adsorbed MCM-41 nanocomposites (samples 1A–C) occurs and UVM-7 nanocomposites (typical of the intra-particle pore
system) practically disappears, exception being made for the
MCM-41 nanocomposite with the lowest complex 1 loading
(sample 1A) (Fig. 5). In this regard, sample 1A constitutes a
particular case where only a portion of the mesopores are filled
and/or blocked by Ni8 molecules. As the complex 1 loading

Fig. 2 Representative TEM micrographs for samples 1E (a), 1G (b), Fig. 4 XRD patterns for samples 1A (a), 1B (b), and 1C (c). The inset
1B (c), and 1I (d). The insets show enlarged (63) TEM images. shows the XRD pattern of the parent MCM-41 pure silica.

2706 | J. Mater. Chem., 2006, 16, 2702–2714 This journal is ß The Royal Society of Chemistry 2006
Fig. 7 Representative N2 adsorption–desorption isotherms for
UVM-11 (a) and 1I (b). The isotherms are vertically shifted for clarity.

1A (628.9 m2 g21) down to the very low value of 1C


(17.6 m2 g21), typical of bulk solids (Table 1).
In the series of Ni8 adsorbed UVM-7 nanocomposites, the
pore volume estimated from partial pressure data at P/P0 , 0.4
Fig. 5 Representative N2 adsorption–desorption isotherms for strongly drops from UVM-7 pure silica to 1D and later
MCM-41 (a), 1A (b), 1B (c), and 1C (d). The isotherms are vertically decreases in a smooth way as the complex 1 content increases
shifted for clarity. for 1E to 1H (Table 1). However, the second adsorption step at
high P/P0 values (typical of the inter-particle textural-like
porosity) is present in all samples (Fig. 6). The pore volume
associated with this last large pore system remains practically
constant for UVM-7 pure silica and the nanocomposites
with low complex loading (samples 1D and 1E), and slowly
decreases as the complex loading increases (samples 1F and
1H) (Table 1). A similar trend is observed for BET surface
areas (Table 1). Samples 1D and 1E present small BET surface
areas (166–146 m2 g21) with a typology of their isotherms
characteristic of UVM-7 materials with filled primary meso-
pores, as observed for mesostructured UVM-7. As the complex
content increases, a small additional surface loss is detected for
1F to 1H (92.5–90.4 m2 g21). By contrast, the small complex 1
loading in the Ni8 adsorbed UVM-11 nanocomposite
(sample 1I) hardly modifies the isotherm, which maintains
the adsorption at high P/P0 pressure values characteristic
of UVM-11 pure silica (Fig. 7). This fact points out the
importance of the mesopore entrances as preferred sites for
host–guest interactions, which explains the inefficiency of the
inter-particle macropore system of the UVM-11 silica surface
for Ni8 adsorption when compared with that of the UVM-7
silica surface.
Overall, these features suggest that complex 1 presents a
Fig. 6 Representative N2 adsorption–desorption isotherms for marked preference to be hosted inside the intra-particle meso-
UVM-7 (a), 1D (b), 1E (c), 1G (d), and 1H (e). The isotherms are pores. Then, the Ni8 molecules partially fill, and consequently
vertically shifted for clarity. block, the intra-particle mesopore system, even for nano-
composites with low complex concentration (samples 1A, 1B,
increases along the series of Ni8 adsorbed MCM-41 nano- and 1D). After the blocking of the intrananoparticle pores, the
composites, there is a sufficient amount of the Ni8 molecule to adsorbed Ni8 molecules must be located at the external particle
partially fill or completely block all the mesopores (samples 1B surface for nanocomposites with high complex concentration
and 1C). The pore volume estimated from partial pressure data (samples 1C and 1E–1H). The relative proportion of Ni8
at P/P0 , 0.4 decreases smoothly as the complex 1 content molecules hosted inside and outside the intra-particle meso-
increases from 1A to 1C (Table 1). The BET surface area also pores, estimated from the evolution of the complex loading
decreases in a progressive way from the relatively high value of amount along each series, depends on the type of mesoporous

This journal is ß The Royal Society of Chemistry 2006 J. Mater. Chem., 2006, 16, 2702–2714 | 2707
host material. Thus, the occupation ratio is 71:29 for the Ni8
adsorbed MCM-41 nanocomposite with the highest complex
concentration (sample 1C), whereas it equals 27:73 for
the corresponding Ni8 adsorbed UVM-7 nanocomposite
(sample 1H). Indeed, this inverted distribution in the relative
occupation of the Ni8 molecules reflects the differences
between the two types of mesoporous host materials, which
are associated with the morphology of the silica support
(unimodal vs. bimodal pore systems) and the pore length
(a few mm vs. 25 nm) and shape (cylindrical vs. worm-like).

Magnetic properties

DC magnetic measurements. The xMT vs. T plots of samples


1A–I, xM being the magnetic susceptibility per Ni8 unit and T
the temperature, are shown in Figs. 8–10. The variation in the
magnetic behavior as the complex loading amount increases
for the two series of Ni8 adsorbed MCM-41 and UVM-7 nano-
composites is consistent with the occurrence of an aggregation
process of Ni8 units in the solid state, regardless of the Fig. 9 Temperature dependence of xMT for 1D (N), 1E (&), 1F (¤),
mesoporous silica nature, either unimodal (samples 1A–C) or 1G (m), and 1H (n) under an applied magnetic field of 1 T (T ¢ 25 K)
bimodal (samples 1D–H) (Figs. 8 and 9, respectively). This is and 250 G (T , 25 K). The inset shows the hysteresis loop of M at
T = 2 K for 1G. The solid lines are eye-guides.
evident when compared with the Ni8 adsorbed UVM-11
nanocomposite (sample 1I), which shows a magnetic behavior
similar to that of complex 1 in the crystalline state (Fig. 10).16
xMT shows a minimum in the temperature range 24–44 K
for 1A–I, which is characteristic of the two antiferromagne-
tically coupled oxamato-bridged S = 2 NiII4 centered triangles
present in 1 (Tmin = 45 K).16 The xMT values at Tmin for 1A–I
(xMTmin = 5.3–5.6 cm3 K mol21) are almost identical to that of
1 (xMTmin = 5.5 cm3 K mol21)15 (Table 2). Yet, the increase of
xMT in the low-temperature region in the two series of Ni8
adsorbed MCM-41 and UVM-7 nanocomposites (Figs. 8
and 9, respectively) is higher than expected for magnetically
isolated S = 4 Ni8 molecules as those present in the UVM-11
nanocomposite (Fig. 10). This behavior supports the presence

Fig. 10 Temperature dependence of xMT for 1I (m) and 1 (n) under


an applied magnetic field of 1 T (T ¢ 25 K) and 250 G (T , 25 K).
The inset shows the hysteresis loops of M at T = 2 K. The solid lines
are eye-guides.

of S = 4 Ni8 repeat units in the resultant [Ni8]x aggregates of


large spin values (S = 4x) for all silica-based mesoporous
nanocomposite materials. Finally, xMT reaches a maximum
in the temperature range 3–10 K for 1A–I, as observed in
1 (Tmax = 3 K),16 which is due to magnetic anisotropy
effects and/or antiferromagnetic interactions between the
aggregates. The xMT values at Tmax increase slowly from 1A
Fig. 8 Temperature dependence of xMT for 1A (N), 1B (&), and 1C to 1C (xMTmax = 6.8–17.8 cm3 K mol21) (Table 2), as the
(¤) under an applied magnetic field of 1 T (T ¢ 25 K) and 250 G complex loading amount increases along this series of Ni8
(T , 25 K). The inset shows the hysteresis loop of M at T = 2 K for adsorbed MCM-41 nanocomposites (Fig. 11). By contrast,
1C. The solid lines are eye-guides. the xMT values at Tmax increase rapidly from 1D to 1G

2708 | J. Mater. Chem., 2006, 16, 2702–2714 This journal is ß The Royal Society of Chemistry 2006
At H = 1 T. At H = 250 G. Calculated from the maxima of xM0 at the different frequencies through the equation t = t0exp[Ueff/kB(T 2 T0)] (with t = 1/2pn and T = Tmax). The values
calculated from the Arrhenius plots (T0 = 0) are given in parentheses. d Calculated from the maxima of xM0 at the different frequencies through the equation k = (DT/T)/Dlog(n) (with T = Tmax).
(xMTmax = 11.3–49.4 cm3 K mol21), and then it decreases

2.70 (9.0)
abruptly for 1H (xMTmax = 11.5 cm3 K mol21) (Table 2), as the

xTe (T/K)

10.3 (9.0)
complex loading amount increases along the series of Ni8
adsorbed UVM-7 nanocomposites (Fig. 11). With only one
exception, the xMT values at Tmax for 1A–H are greater than
that of 1I (xMTmax = 8.4 cm3 K mol21) (Table 2), which in turn

0.08 (9.0)
xSe (T/K)

0.3 (9.0)
is almost identical to that of 1 (xMTmax = 8.6 cm3 K mol21)16
(Fig. 11). This behavior indicates the occurrence of an inter-
molecular ferromagnetic coupling between the S = 4 Ni8

Calculated from the fit of the xM0 vs. xM9 plots through the equation xM0 = {( xT 2 xS)/2tan[(1 + 3)p/2]}{(xM9 2 xS)( xT 2 xM9) + ( xT 2 xS)2/4tan2[(1 + 3)p/2]}1/2.
0.81 (9.0)

0.76 (9.0)
units in the resultant aggregate for 1A–G, with the onset of a
ae (T/K)

relatively strong antiferromagnetic interaction between the


aggregates for 1H.
From the available physical and magnetic characterization
0.031
0.028
0.014
0.017
0.009
0.014
0.019

0.014
data, some preliminary considerations can be made on the
kd

mechanism for the nucleation and growth of the aggregates of


Ni8 molecules adsorbed on the mesoporous silica surface, as
T0c/K

shown in Scheme 2. The adsorptive behavior shows that the


1.6
1.7
6.3
2.9
5.5
5.7
5.8

1.9

initial incorporation of individual Ni8 molecules takes place


inside the intra-particle mesopores which become the preferen-
10214)
10215)
10232)
10227)
10250)
10234)
10224)

0.9 6 10211 (1.0 6 10220)

tial sites for surface adsorption due to strong capillarity effects.


More likely, the isolated Ni8 molecules are bonded to the
6
6
6
6
6
6
6

mesopore walls by anchoring silanolate groups which counter-


(1.3
(0.6
(0.5
(1.7
(1.3
(1.8
(2.0

balance the cationic charge (perchlorate is absent in all of the


10211
10211
10211
10211
10211
10211
10211

studied materials) (Scheme 2(a)). After the filling of the intra-


particle mesopores, the Ni8 molecules that are located close
6
6
6
6
6
6
6
t0c/s

to the mesopore entrances start the growing of small [Ni8]x


3.0
1.5
0.5
2.0
1.5
1.0
1.0

aggregates at the external inter-particle surface (Scheme 2(b)).


The importance of these sites as nucleation centres is
Ueffc/cm21

40 (100)
42 (435)
20 (170)
21 (520)
35 (415)
55 (290)

30 (120)
33 (95)

confirmed by the absence of Ni8 aggregation in the UVM-11


material because of the lack of intra-particle porosity. The
poor adsorptive capability and the magnetic phenomenology
of this last material show that the interactions of Ni8 molecules
TB/K (n/Hz)

(1000)
(1000)
(1000)
(1000)
(1000)
(1000)
(1000)
(1000)
(1000)

with a non-porous silica surface are relatively weak and labile.


On the other hand, the high degree of aggregation estimated
10.5
5.0
5.3
9.5
4.6
7.0
8.6

,1.8
4.5

from the magnetic data in the series of Ni8 adsorbed UVM-7


nanocomposites (samples 1D–G) benefits from the small intra-
particle mesopore length and the large inter-particle porosity
Hc/G (T/K)

(2.0)
(2.0)
(2.0)
(2.0)
(2.0)
(2.0)
(2.0)
(2.0)
(2.0)
2990
140
325
800
140
460
790

240
65
xMTmaxb/cm3 K mol21

(10.0)
(6.5)
(5.5)
(9.0)
(5.3)
(7.0)
(7.5)

(3.8)
(6.5)
(Tmax/K)
Table 2 Selected magnetic data for 1A–I

12.4
17.8
11.3
15.6
31.5
49.4
11.5
6.8

8.4
c
xMTmina/cm3 K mol21

(24.0)
(28.0)
(30.0)
(30.0)
(34.0)
(30.0)
(44.0)
(30.0)
(30.0)
(Tmin/K)

b
5.5
5.6
5.3
5.5
5.4
5.4
5.4
5.6
5.3

Fig. 11 Dependence of the xMT values at Tmax for 1A–I on the initial
Sample

relative concentration of 1 (data from Table 2). The horizontal line


1H
1D

1G
1A

1C
1B

1E
1F

1I

corresponds to the xMT value at Tmax for 1.


a

This journal is ß The Royal Society of Chemistry 2006 J. Mater. Chem., 2006, 16, 2702–2714 | 2709
Scheme 2 Proposed mechanism for the incorporation and aggregation of cationic Ni8 cage complex 1 into mesoporous silica: (a) intra-
particle adsorption at the mesopore walls by anchoring silanolate groups, (b) and (c) inter-particle adsorption at the external surface by linking
silanolate groups.

when compared with the Ni8 adsorbed MCM-41 ones The M vs. H plots of 1A–I at 2.0 K, M being the
(samples 1A–C). In fact, the aggregation on the external magnetization per Ni8 unit and H the applied magnetic
inter-particle surface is presumably due to the mediated field, show hysteresis loops which are indicative of slow
effect of the deprotonated silanol groups that allows the Ni8 magnetic relaxation effects (inset of Figs. 8–10). Moreover, the
condensation through Si–O bridges. The inter-molecular coercivities vary differently as the complex loading amount
ferromagnetic exchange interactions are then propagated by increases along each series of silica-based mesoporous
the silanolate bridges between the peripheral high-spin NiII materials (Fig. 12). In the series of Ni8 adsorbed MCM-41
ions of neighboring Ni8 molecules. nanocomposites, the coercive field values increase con-
The growing of the large [Ni8]x aggregates, probably of tinuously as the complex loading amount increases from 1A
filamentous-like shape because of the interacting topologies to 1C (Hc = 140–800 G) (Table 2). Similarly, as the complex
of host surface and guest species, stops when discrete loading amount increases in the series of Ni8 adsorbed UVM-7
neighbouring aggregates meet each other and condense
through additional silanolate bridges. The resultant ‘super-
aggregate’ should be consistent with both the C3h molecular
symmetry of the Ni8 units and the hexagonal order of the
mesoporous silica (Scheme 2(c)). The occurrence of such a
self-organization process with increasing complex aggregation
should be favored by the multi-domain surface morphology
and irregular topography of the large cage-like inter-particle
pore system of UVM-7 silica. In this case, the steric
requirements associated with this process dictate a change in
the nature and/or the magnitude of the inter-molecular
interactions through the silanolate bridges from weakly
ferromagnetic within the aggregates to strongly antiferro-
magnetic between the aggregates. This is ultimately responsible
for the ‘magnetic collapse’ observed in the Ni8 adsorbed
UVM-7 nanocomposite with the highest complex loading
amount (sample 1H). In fact, it is well-known that the nature
of the magnetic coupling in related m-hydroxo, m-alkoxo,
and m-phenoxo nickel(II) complexes depends critically on the
Ni–O–Ni bridge angle (c), passing from ferromagnetic to Fig. 12 Dependence of the Hc values at T = 2 K for 1A–I on the
antiferromagnetic when increasing the c value (for c values initial relative concentration of 1 (data from Table 2). The horizontal
close to 90u).19,20 line corresponds to the Hc value at T = 2 K for 1.

2710 | J. Mater. Chem., 2006, 16, 2702–2714 This journal is ß The Royal Society of Chemistry 2006
nanocomposites, the coercive field first increases from 1D to AC magnetic measurements. The xM9 and xM0 vs. T plots
1G, reaching extremely high values (Hc = 140–2990 G), but for 1A–I, xM9 and xM0 being the in-phase and out-of-phase
then it drops dramatically for 1H (Hc = 65 G) (Table 2). magnetic susceptibility per Ni8 unit under zero DC field,
Interestingly, the coercive field value for 1I (Hc = 240 G) respectively, exhibit a frequency-dependent magnetic behavior,
(Table 2), which contains magnetically isolated S = 4 Ni8 as expected for SMMs (Fig. 13). Indeed, frequency-dependent
molecules adsorbed at the UVM-11 silica surface, is even xM9 and xM0 maxima are seen for 1A–G and 1I in the
greater than that of 1 in the crystalline state, which is almost temperature range 4.5–10.5 K, but no maxima are observed
non-negligible (Hc = 25 G). for 1H above 1.8 K (Fig. S1, ESI{). As the frequency of the
1 G AC field increases in the range 1–1000 Hz, the maxima
of both xM9 and xM0 for 1A–G and 1I are shifted toward
higher temperatures. Interestingly, the xM9 and xM0 maxima
for 1A–G occur at higher temperatures than those of 1I
and, more importantly, they shift to higher temperatures on
progressive aggregation along both series of mesoporous silica-
based nanocomposite materials.
In fact, the temperature value of the maximum of xM0 at a
given frequency (n) corresponds to the blocking temperature
(TB = Tmax), whereby it is assumed that the switching of
the oscillating AC field matches the relaxation rate of the
magnetization (1/t = 2pn). Thus, as the complex loading
amount increases in the series of Ni8 adsorbed UVM-7
nanocomposites, the blocking temperature value at n =
1000 Hz first remains practically constant from 1A to 1B
(TB = 5.0–5.3 K) and then it suddenly increases for 1C (TB =
9.5 K) (Table 2). By contrast, they increase steadily for 1D–G
(TB = 4.6–10.5 K) (Table 2) as the complex loading amount
increases in the series of Ni8 adsorbed UVM-7 nanocompo-
sites. The different evolution of the blocking temperature
values between the two types of mesoporous host materials
(Fig. 14) reflects the differences in the relative proportion of
Ni8 molecules hosted inside and outside the intra-particle
mesopores along each series, as discussed above. Thus, the
higher TB values for the nanocomposites with high complex
concentration (samples 1C and 1E–1G) are associated with the
presence of [Ni8]x aggregates located at the external particle
surface (Scheme 2(b)). They will relax more slowly than the
isolated Ni8 molecules within the intra-particle mesopores
in the nanocomposites with low complex concentration
(samples 1A, 1B, and 1D) (Scheme 2(a)). In this regard, the
nanocomposite 1G, which is obtained from the ferromagnetic

Fig. 13 Temperature dependence of xM9 and xM0 (open symbols) for


1C (a), 1G (b), and 1I (c) in zero DC field and under 1G AC field at Fig. 14 Dependence of the TB values at n = 1000 Hz for 1A–G and 1I
oscillating frequencies of 1 (N), 10 (&), 100 (¤), and 1000 Hz (m). The on the initial relative concentration of 1 (data from Table 2). The
solid lines are eye-guides. horizontal line corresponds to the TB value at n = 1000 Hz for 1.

This journal is ß The Royal Society of Chemistry 2006 J. Mater. Chem., 2006, 16, 2702–2714 | 2711
arrangement of Ni8 units at the inter-particle pores of UVM-7
bimodal mesoporous silica, possesses the highest blocking
temperature yet reported for well-established nickel SMMs.21
On the other hand, the nanocomposite 1I which contains
magnetically isolated S = 4 Ni8 molecules adsorbed at the
UVM-11 silica surface possesses a higher blocking temperature
than that of 1 in the crystalline state. Thus, the TB value at
n = 1000 Hz for 1I is 4.5 K (Table 2), whereas it is 3.0 K for 1.16
This is likely due to the larger magnetic anisotropy (D) of the
adsorbed Ni8 molecules because of the structural modifications
that would accompany the process of incorporation within
the silica host.
The calculated relaxation time (t = 1/2pn) at Tmax for 1A–G
and 1I follows the Arrhenius law characteristic of a thermally-
activated mechanism, t = t0exp(Ueff/kBT) (Fig. S2, ESI{). Yet,
the values of the effective energy barrier for the reversal of the Fig. 15 Dependence of the T0 values for 1A–G and 1I on the initial
magnetization are very high (Ueff = 95–520 cm21) and the pre- relative concentration of 1 (data from Table 2).
exponential factor values are abnormally low (t0 = 1.3 6
10214–1.3 6 10250 s) (Table 2). These anomalous Arrhenius- 1D–G) (Table 2). In general, the magnitude of the effective
fitting parameters are reminiscent of a classical spin glass energy barrier (Ueff = 2DS2) depends on a number of factors,
system for which the relaxation time follows an Arrhenius-like including not only S, which increases monotonically with the
behavior but with physical meaningless values for the energy degree of aggregation, but also D, which would vary in an
barrier and the pre-exponential factor.22 In fact, the values unpredictable manner from one aggregate to another accord-
of the relative variation of Tmax per decade of frequency, k = ing to the structural modifications (molecular symmetry
(DT/T)/Dlog(n), for 1A–G and 1I are in the range 0.01–0.03 lowering) that would accompany the process of incorporation
(Table 2).22,23 These very low k values place these Ni8 adsorbed and aggregation in each mesoporous silica support.
silica-based nanocomposites between typical spin glasses and In our case, we suggest that the spin glass behavior stems
superparamagnets, suggesting a picture of weakly interacting from the existence of a wide distribution of [Ni8]x aggregates
clusters (‘cluster glasses’).22 with different conformation and association degree (size
That being so, the thermal variation of the relaxation time at distribution) and the presence of weak interactions between
Tmax for 1A–G and 1I has been fitted through the Vogel– the aggregates as well. These two factors contribute to the
Fulcher law for interacting systems, t = t0exp[Ueff/kB(T 2 T0)] different energy barriers and, consequently, different relaxa-
(Fig. S2, ESI{).22,24 Thus, the calculated T0 values in the range tion times that will lead to the observed non-Arrhenius law
1.6–6.3 K give an estimation of the magnitude of the inter- behavior for this family of Ni8 adsorbed silica-based nano-
cluster interaction in these Ni8 adsorbed silica-based nano- composites. Thus, for instance, the xM0 vs. xM9 plots (so-called
composites (Table 2). As was shown earlier for the blocking Cole–Cole plots) of 1C and 1G at 9.0 K in the frequency range
temperature, the T0 values increase for 1A–C (T0 = 1.6–6.3 K) 0.1–1400 Hz yield rather flattened semicircles (Fig. 17), which
and 1D–G (T0 = 2.9–5.8 K) (Table 2), in measure with the supports the occurrence of a wide distribution of single
increase in the degree of aggregation along both series of Ni8 relaxation processes. This is confirmed by the fit of the
adsorbed MCM-41 and UVM-7 nanocomposites, respectively experimental data through a generalized Debye model, which
(Fig. 15). This is as expected for dipole–dipole interactions gives very large values for the a parameter of 0.81 for 1C and
between the [Ni8]x aggregates, which increase with the spin
value (S = 4x) and the close packing of the aggregates. The fact
that the T0 value for 1I is small but non-negligible (T0 = 1.9 K)
(Table 2) may also suggest that the S = 4 Ni8 molecules
adsorbed at the UVM-11 silica surface are not perfectly iso-
lated but weakly interact through dipole–dipole interactions.
On the other hand, the calculated values of the effective
energy barrier (Ueff = 20–55 cm21) and the pre-exponential
factor (t0 = 0.5 6 10211–3.0 6 10211 s) for 1A–G and 1I
through the Vogel–Fulcher law (Table 2) are now physically
meaningful and in agreement with those obtained in other
well-established SMMs. Notably, the Ueff values for 1A–G
increase on progressive aggregation along both series of silica-
based nanocomposite materials (Fig. 16). However, the
relative variation in the effective energy barrier is somewhat
smaller along the series of Ni8 adsorbed MCM-41 nano-
composites (Ueff = 33–42 cm21 for 1A–C) when compared with Fig. 16 Dependence of the Ueff values for 1A–G and 1I on the initial
the series of Ni8 adsorbed UVM-7 ones (Ueff = 20–55 cm21 for relative concentration of 1 (data from Table 2).

2712 | J. Mater. Chem., 2006, 16, 2702–2714 This journal is ß The Royal Society of Chemistry 2006
M. A. Ratner, P. J. Rossky, S. I. Stupp and M. E. Thompson,
Adv. Mater., 1998, 10, 1297; E. Dujardin and S. Mann, Adv.
Mater., 2004, 16, 1125.
2 Thematic issues on nanoscopic functional materials: Chem.
Mater., 2001, 13, 3059; Chem. Rev., 2005, 105, 1023.
3 C. T. Kresge, M. E. Leonowicz, W. J. Roth, J. C. Vartuli and
J. S. Beck, Nature, 1992, 359, 710.
4 Special issue on novel porous materials for emerging applications:
J. Mater. Chem., 2006, 16.
5 (a) M. Comes, G. Rodrı́guez-López, M. D. Marcos, R. Martı́nez-
Máñez, F. Sancenón, J. Soto, L. A. Villaescusa, P. Amorós and
D. Beltrán, Angew. Chem., Int. Ed., 2005, 44, 2918; (b)
A. B. Descalzo, K. Rurack, H. Weißhoff, R. Martı́nez-Máñez,
M. D. Marcos, P. Amorós and J. Soto, J. Am. Chem. Soc., 2005,
127, 184; (c) M. Comes, M. D. Marcos, R. Martı́nez-Máñez,
F. Sancenón, J. Soto, L. A. Villaescusa, P. Amorós and D. Beltrán,
Adv. Mater., 2004, 16, 1783.
6 (a) R. Sessoli, H. L. Tsai, A. R. Schake, S. Wang, J. B. Vincent,
K. Folting, D. Gatteschi, G. Christou and D. N. Hendrickson,
J. Am. Chem. Soc., 1993, 115, 1804; (b) R. Sessoli, D. Gatteschi,
A. Caneschi and M. A. Novak, Nature, 1993, 365, 141.
7 D. Gatteschi and R. Sessoli, Angew. Chem., Int. Ed., 2003, 42, 268.
8 (a) M. Soler, W. Wernsdorfer, K. Folting, M. Pink and
G. Christou, J. Am. Chem. Soc., 2004, 126, 2156; (b)
A. Tasiopoulos, A. Vinslava, W. Wernsdorfer, K. A. Abboud
Fig. 17 Cole–Cole plots for 1C (n) and 1G (m) at T = 9 K in zero and G. Christou, Angew. Chem., Int. Ed., 2004, 43, 2117.
DC field and 1G AC field at oscillating frequencies in the range 9 (a) M. Clemente-León, H. Soyer, E. Coronado, C. Mingotaud,
C. J. Gómez-Garcı́a and P. Delhaes, Angew. Chem., Int. Ed., 1998,
0.1–1400 Hz. The solid lines are the best-fit curves.
37, 2842; (b) M. Clemente-León, E. Coronado, A. Forment-Aliaga,
J. M. Martı́nez-Agudo and P. Amorós, Polyhedron, 2003, 22, 2395;
0.76 for 1G (a = 0 for a single relaxation process, while a = 1 (c) M. Clemente-León, E. Coronado, A. Forment-Aliaga,
P. Amorós, J. Ramı́rez-Castellanos and J. M. González-Calbet,
for an infinite distribution of single relaxation processes). The J. Mater. Chem., 2003, 13, 3089; (d) E. Coronado, A. Forment-
adiabatic (xS) and isothermal (xT) susceptibility values are Aliaga, F. M. Romero, V. Corradini, R. Biagi, V. De Renzi,
0.08 and 2.70 cm3 mol21 for 1C, and 0.30 and 10.3 cm3 mol21 A. Gambardella and U. Del Pennino, Inorg. Chem., 2005, 44, 7693.
10 (a) T. Coradin, J. Larionova, A. A. Smith, G. Rogez, R. Clérac,
for 1G, respectively.22,25 C. Guérin, G. Blondin, R. E. P. Winpenny, C. Sanchez and
T. Mallah, Adv. Mater., 2002, 14, 896; (b) S. Willemin,
Conclusions G. Arrachart, L. Lecren, J. Larionova, T. Coradin, R. Clérac,
T. Mallah, C. Guérin and C. Sanchez, New J. Chem., 2003, 27,
1533; (c) B. Folch, J. Larionova, Y. Guari, C. Guérin, A. Mehdi
In this work, we have demonstrated that ordered mesoporous
and C. Reyé, J. Mater. Chem., 2004, 14, 2703.
silicas provide appropriate template conditions for the surface 11 (a) D. Ruiz-Molina, M. Mas-Torrent, J. Gómez, A. I. Balana,
(‘heterogeneous’) aggregation of suitably designed small N. Domingo, J. Tejada, M. T. Martı́nez, C. Rovira and J. Veciana,
octanuclear nickel(II) SMMs with a relatively large ground Adv. Mater., 2003, 15, 42; (b) M. Cavallini, F. Biscarini, J. Gómez-
Segura, D. Ruiz and J. Veciana, Nano Lett., 2003, 3, 1527; (c)
spin state and a significant magnetic anisotropy. The resultant M. Cavallini, J. Gómez-Segura, D. Ruiz-Molina, M. Massi,
complex aggregates obtained from the ferromagnetic C. Albonetti, C. Rovira, J. Veciana and F. Biscarini, Angew.
arrangement of Ni8 complex precursors show slow relaxation Chem., Int. Ed., 2005, 44, 888; (d) J. Gómez-Segura, O. Kazakova,
of the magnetization at higher blocking temperatures and J. Davis, P. Josephs-Franks, J. Veciana and D. Ruiz-Molina,
Chem. Commun., 2005, 5615.
unique spin glass behavior. This new synthetic approach 12 (a) A. Cornia, A. C. Fabretti, M. Pacchioni, L. Zobbi, B. Bonacchi,
in confined mesoporous media may be an attractive alternative A. Caneschi, D. Gatteschi, R. Biagi, U. Del Pennino, V. De Renzi,
to solution (‘homogeneous’) aggregation methods for the L. Gurevich and H. S. J. Van der Zant, Angew. Chem., Int. Ed.,
2003, 42, 1645; (b) G. C. Condorelli, A. Motta, I. L. Fragala,
development of high-TB slow relaxing magnetic nanocompo- F. Giannazzo, V. Ranieri, A. Caneschi and D. Gatteschi, Angew.
site materials. Chem., Int. Ed., 2004, 43, 4081.
13 A. N. Abdi, J. P. Bucher, P. Rabu, O. Toulemonde, M. Drillon and
P. Gerbier, J. Appl. Phys., 2004, 95, 7345.
Acknowledgements 14 J. S. Steckel, N. S. Persky, C. R. Martı́nez, C. L. Barnes, E. A. Fry,
J. Kulkarni, J. D. Burgess, R. B. Pacheco and S. L. Stoll, Nano
This work was supported by the Spanish Ministry of Science Lett., 2004, 4, 399.
and Technology (Projects MAT2002-04329-C03-02 and 15 F. Palacio, P. Oliete, U. Schubert, I. Mijatovic, N. Hüsing and
MAT2003-08568-C03-01), the Generalitat Valenciana H. Peterlik, J. Mater. Chem., 2004, 14, 1873.
16 E. Pardo, I. Morales-Osorio, M. Julve, F. Lloret, J. Cano, R. Ruiz-
(GRUPOS03/099), and the European Union through the
Garcı́a, J. Pasán, C. Ruiz-Pérez, X. Ottenwaelder and Y. Journaux,
Magmanet Network of Excellence (Contract 515767-2). P. B. Inorg. Chem., 2004, 43, 7594.
and E. P. thank the Spanish Ministry of Education, Culture, 17 (a) E. Pardo, K. Bernot, M. Julve, F. Lloret, J. Cano, R. Ruiz-
and Sports for a grant. Garcı́a, F. S. Delgado, C. Ruiz-Pérez, X. Ottenwaelder and
Y. Journaux, Inorg. Chem., 2004, 43, 2768; (b) E. Pardo,
K. Bernot, M. Julve, F. Lloret, J. Cano, R. Ruiz-Garcı́a,
References J. Pasán, C. Ruiz-Pérez, X. Ottenwaelder and Y. Journaux,
Chem. Commun., 2004, 920; (c) X. Ottenwaelder, J. Cano,
1 A. P. Alivisatos, P. F. Barbara, A. W. Castleman, J. Chang, Y. Journaux, E. Rivière, C. Brennan, M. Nierlich and R. Ruiz-
D. A. Dixon, M. L. Klein, G. L. McLendon, J. S. Miller, Garcı́a, Angew. Chem., Int. Ed., 2004, 43, 850.

This journal is ß The Royal Society of Chemistry 2006 J. Mater. Chem., 2006, 16, 2702–2714 | 2713
18 (a) S. Cabrera, J. El Haskouri, C. Guillem, J. Latorre, A. Beltrán, C. M. Grant, H. Gudel, M. Murrie, S. Parsons, C. Paulsen,
D. Beltrán, M. D. Marcos and P. Amorós, Solid State Sci., 2000, 2, F. Semadini, V. Villar, W. Wernsdorfer and R. E. P. Winpenny,
405; (b) P. Amorós, A. Beltrán, D. Beltrán, S. Cabrera, J. El Chem.–Eur. J., 2002, 8, 4867; (c) S. T. Ochsenbein, M. Murrie,
Haskouri and M. D. Marcos, Patent WO 01-72635; (c) J. El E. Rusanov, H. Stœckli-Evans, C. Sekine and H. U. Güdel,
Haskouri, D. Ortiz de Zárate, C. Guillem, J. Latorre, M. Caldés, Inorg. Chem., 2002, 41, 5133; (d) R. S. Edwards, S. Maccagnano,
A. Beltrán, D. Beltrán, A. B. Descalzo, G. Rodrı́guez-López, E. C. Yang, S. Hill, W. Wernsdorfer, D. N. Hendrickson and
R. Martı́nez-Máñez, M. D. Marcos and P. Amorós, Chem. G. Christou, J. Appl. Phys., 2003, 93, 7807; (e) E. C. Yang,
Commun., 2002, 330. W. Wernsdorfer, S. Hill, R. S. Edwards, M. Nakano,
19 (a) J. E. Andrew and A. B. Blake, J. Chem. Soc. A, 1969, 1456; (b) S. Maccagnano, L. N. Zakharov, A. L. Rheingold, G. Christou
J. A. Bertrand, A. P. Ginsberg, R. I. Kaplan, C. E. Kirkwood, and D. N. Hendrickson, Polyhedron, 2003, 22, 1727; (f)
R. L. Martin and R. C. Sherwood, Inorg. Chem., 1971, 10, 240; (c) E. C. Yang, W. Wernsdorfer, L. N. Zakharov, Y. Karaki,
W. L. Gladfelter, M. W. Lynch, W. P. Schaefer, D. N. Hendrickson A. Yamaguchi, R. M. Isidro, G. D. Lu, S. A. Wilson,
and H. B. Gray, Inorg. Chem., 1981, 20, 2390; (d) L. Ballester, A. L. Rheingold, H. Ishimoto and D. N. Hendrickson, Inorg.
E. Coronado, A. Gutiérrez, A. Monge, M. F. Perpiñán, E. Pinilla Chem., 2006, 45, 529.
and T. Rico, Inorg. Chem., 1992, 31, 2053. 22 J. A. Mydosh, Spin Glasses: An Experimental Introduction, Taylor
20 (a) K. K. Nanda, L. K. Thompson, J. N. Bridson and K. Nag, and Francis, London, 1993.
J. Chem. Soc., Chem. Commun., 1994, 1337; (b) K. K. Nanda, 23 M. A. Gı̂rtu, J. Optoelectron. Adv. Mater., 2002, 4, 85.
R. Das, L. K. Thompson, K. Venkastsubramanian, P. Paul and 24 X. X. Zhang, G. H. Wen, G. Xiao and S. Sun, J. Magn. Magn.
K. Nag, Inorg. Chem., 1994, 33, 1188. Mater., 2003, 261, 21.
21 (a) C. Cadiou, M. Murrie, C. Paulsen, V. Villar, W. Wernsdorfer 25 (a) K. S. Cole and R. H. Cole, J. Chem. Phys., 1941, 9, 341; (b)
and R. E. P. Winpenny, Chem. Commun., 2001, 2666; (b) C. J. F. Boettcher, Theory of Electric Polarisation, Elsevier,
H. Andres, R. Basler, A. J. Blake, C. Cadiou, G. Chaboussant, Amsterdam, 1952.

2714 | J. Mater. Chem., 2006, 16, 2702–2714 This journal is ß The Royal Society of Chemistry 2006

Вам также может понравиться