You are on page 1of 5

Journal of

Materials Chemistry C
View Article Online

Published on 07 April 2015. Downloaded by Johannes Gutenberg Universitaet Mainz on 17/08/2015 15:25:06.

PAPER

Cite this: J. Mater. Chem. C, 2015,


3, 4737

View Journal | View Issue

Chiral spin crossover nanoparticles and gels with


switchable circular dichroism
Ilya A. Guralskiy,*a Viktor A. Reshetnikov,a Agnieszka Szebesczyk,b
Elzbieta Gumienna-Kontecka,b Andriy I. Marynin,c Sergii I. Shylin,ad
Vadim Ksenofontovd and Igor O. Fritskya
Spin crossover complexes represent spectacular examples of molecular switchable materials. We describe a
new approach towards homochiral coordination nanoparticles of [Fe(NH2trz)3](L-CSA)2 (NH2trz = 4-amino1,2,4-triazole, L-CSA = L-camphorsulfonate) that display an abrupt switch of chiral properties associated with
a cooperative spin transition. This is an original method that generates stable and additive-free colloidal

Received 17th January 2015,


Accepted 2nd April 2015

solutions of nanoparticles with a spin transition around room temperature. The introduction of a chiral anion

DOI: 10.1039/c5tc00161g

properties that are dierent in high and low spin states. We also illustrate here an eect of the chiral

www.rsc.org/MaterialsC

memory which is caused by the hysteretic spin transition. In addition, a bistable chiral composite
supramolecular system a [Fe(NH2trz)3](L-CSA)2CHCl3 gel is described here.

to the coordination framework makes these nanoparticles display specific chiro-optical (circular dichroism)

Introduction
Chirality, which is an important phenomenon in biology,
physics and chemistry, plays a crucial role in many photonic
experiments.1 This eect of molecular asymmetry is well studied
for organic and biological materials, whereas it is less reviewed
for inorganic and coordination materials, for nanoobjects.2
Nevertheless, a strong chiro-optical response from the interaction of light with chiral plasmonic nanostructures is attracting
much attention. Such nanostructures can be used to enhance
the chiro-optical signal of chiral molecules and can also significantly increase the enantiomeric excess of direct asymmetric
synthesis and catalysis.3 These eects oer numerous implications for biological systems.4 Magnetic chiral colloids that
can rapidly self-assemble into a variety of isomeric forms have
been proposed.5 Various chiral nanoparticles of inorganic semiconductors can be obtained with diverse characteristics of chirooptical activity and excitonic resonance.6
Chiral switches that can be triggered with dierent external
stimuli oer fascinating opportunities as elements in molecular
a

Department of Chemistry, Taras Shevchenko National University of Kyiv,


Volodymyrska St. 64, Kyiv 01601, Ukraine. E-mail: illia.guralskyi@univ.kiev.ua
b
Faculty of Chemistry, University of Wroclaw, F. Joliot-Curie 14, Wroclaw 50383,
Poland
c
Problem Research Laboratory, National University of Food Technologies,
Volodymyrska St. 68, Kyiv 01601, Ukraine
d
Institute of Inorganic and Analytical Chemistry, Johannes Gutenberg University of
Mainz, Staudingerweg 9, 55099 Mainz, Germany
Electronic supplementary information (ESI) available. See DOI: 10.1039/
c5tc00161g

This journal is The Royal Society of Chemistry 2015

devices and photonic materials.7 Thus, a simple yet elegant example


of a chiral bistable system is the one reported recently by Reichert
and Breit,8 which employs the axial chirality of bisphenol and
the conformational chemistry of a six-membered ring.
Among various bistable materials, the spin crossover (SCO)
complexes form a large group of thermo-, photo-, magneto-,
piezo- and chemo-switchable compounds.913 Recent remarkable achievements in studies of SCO allowed their implementation into electronic circuits,14 MEMS,15 micro-thermometers,16
chemical sensors,1721 etc. Photonic applications should be
particularly significant, taking into account that all optical
properties of a SCO object are strongly related to the electronic
states of its metallic centres.22
Elaboration studies of various SCO nanoparticles have led to
impressive developments in the chemistry of spin transition
complexes in the last 10 years.23 Water-in-oil reverse micelles
have been actively applied to elaborate nanoparticles from the
Fe(II)triazole2429 and Hofmann clathrate3032 families. These works
demonstrated the conservation of a first order phase transition
in coordination nanoparticles down to several nanometers in
size. Some of these results showed size effects such as the loss
of cooperativity and the change of transition temperatures.
Chiral SCO frameworks could be obtained by separating
optically active crystals;33,34 however, this method cannot be
applied to produce higher quantities of chiral complexes and is
not suitable for reaching non-monocrystalline forms of homochiral complexes. We describe here the generation of the necessary
asymmetry in final SCO complex materials by using chiral
precursors; e.g., a chiral anion may be introduced to obtain a
chiral coordination framework. In this paper we describe a

J. Mater. Chem. C, 2015, 3, 4737--4741 | 4737

View Article Online

Paper

Journal of Materials Chemistry C

pathway towards homochiral SCO nanomaterials and show


how the cooperative process of an electronic switch aects
the chiro-optical properties of coordination nanoparticles.

Published on 07 April 2015. Downloaded by Johannes Gutenberg Universitaet Mainz on 17/08/2015 15:25:06.

Results and discussion


Chiral nanoparticles
Surfactant-free approaches towards the generation of SCO nanoparticles are of great interest with regard to their future processing
and implementation.35 In this work we used a homogeneous
synthesis of SCO nanoparticles in acetonitrile. The nanoparticles of
poly[tris-(4-amino-1,2,4-triazolato)iron(II) camphorsulfonate] (Fig. 1a)
were obtained as a stable colloid (1) in MeCN by a reaction between
the corresponding iron salt and the ligand. In this case acetonitrile
plays the role both of a solvent and a stabilizer that prevents the
nanoparticles from aggregating. The latter eect should be mostly
related to the coordination of MeCN molecules to the surface Fe(II)
ions with an unsaturated coordination environment.
Dynamic light scattering (DLS), which is a powerful approach
for monitoring nanoparticles in solutions, shows a typical distribution of nanoparticles in colloid 1 (Fig. 1b) with a size of 79 
22 nm. The TEM measurements (Fig. 1c and d, Fig. S1 and S2,
ESI) indicate that colloid 1 forms slightly elongated nanoparticles with a longer dimension of about (80  17) nm and a
form factor of about 2.3. The size of the nanoparticles can be
slightly tuned by changing the concentration of the precursors
(Fig. S3, ESI). Both the powder X-ray diraction and the electron
diraction patterns for individual nanoparticles confirm their
almost amorphous structure (Fig. S4, ESI).
Temperature dependent UV-Vis measurements were taken for
colloid 1. Although a concentrated solution of nanoparticles is
highly turbid and unsuitable for these kinds of measurements,
the absorption by a diluted solution can be easily followed.

Fig. 1 Structure and morphology of nanoparticles 1. (a) Schematic representation of a typical Fe(II) 1,2,4-triazole complex structure. Fe(II) ions
are known to form 1D-chains with 4-amino-1,2,4-triazole,36 whereas
anions are situated in the inter-chain channels. (b) DLS measurements
on the colloidal solution of nanoparticles 1. Size distribution is given as
the percent of nanoparticles (distribution by number) from 3 dierent
measurements (298 K). (c) TEM image of nanoparticles 1. (d) Distribution of
nanoparticles by size from the TEM image.

4738 | J. Mater. Chem. C, 2015, 3, 4737--4741

Fig. 2 Cooperative spin transition in nanoparticles of [Fe(NH2trz)3](L-CSA)2.


(a) Temperature dependent UV-Vis spectra (240380 nm) at 304329 K of a
colloidal solution of nanoparticles 1 in acetonitrile. (b) Temperature induced
change of the MLCT band at 280 nm. (c) The SQUID measurement on dry
nanoparticles. The wMT vs. T dependence demonstrates a cooperative transition between the diamagnetic and paramagnetic states of SCO nanoparticles
(T1/2 = 316 Km and 300 Kk). The Mo
ssbauer spectra in two spin states for
these nanoparticles are inserted.

Temperature dependent spectra of 1 are shown in Fig. 2a.


The band that is characteristic for the LS form is related to the
metal-to-ligand charge transfer (MLCT) and it decreases with
temperature due to the conversion to the HS form. The high
extinction coecient of the MLCT band (e = 30 000 L mol 1 cm 1)
permitted the measurements even for a much diluted solution
(10 6 mol L 1). By tracing the temperature dependent changes
of e, the spin transition in the studied nanoparticles could be
followed (Fig. 2b). The curve indicates a cooperative spin transition that takes place at 318 K in heating mode and at 309 K in
cooling mode. Both the abruptness of the SCO and the hysteresis
loop demonstrate the cooperativity of the spin transition.
Removal of solvent molecules from nanoparticles of 1 leads
to a typical solvent eect that is usually observed in this family
of complexes.37 As can be seen from the wMT vs. T plot (Fig. 2c,
Fig. S5, ESI) the temperature of transition slightly decreases:
T1/2 is 316 K and 300 K in heating and cooling modes,
respectively. In the 4280 K range and above 325 K, the wMT values
stay constant and are typical for LS and HS Fe(II) complexes.
This transition is complete in both directions as confirmed by
ssbauer measurements. Typical doublets are observed in the LS
Mo
(d = 0.43 mm s 1, D = 0.21 mm s 1) and the HS (d = 0.99 mm s 1,
D = 2.63 mm s 1) states. Importantly, the spin transition in
nanoparticles 1 can take place both in the solid state and in
solution which enriches the possibilities for their application.
Circular dichroism (CD) of an optically active material
describes the dierential absorption of left- and right-handed
circularly polarized light. It has a wide range of applications in
many dierent fields, most notably, to investigate the secondary
structure of proteins. For coordination compounds, CD spectroscopy appears to be a preferential approach in investigating
chiral complexes. A CD spectrum can only be measured for a

This journal is The Royal Society of Chemistry 2015

View Article Online

Published on 07 April 2015. Downloaded by Johannes Gutenberg Universitaet Mainz on 17/08/2015 15:25:06.

Journal of Materials Chemistry C

Paper

structure of the complex; (b) changes of chirality related to the


spatial distortions that accompany the spin transition. Due to
the hysteresis of the spin transition, the chiro-optical behaviour
of nanoparticles 1 is dierent in heating and cooling modes
causing a chiral memory eect (in this case the ability
to possess one of the two chiral states at given conditions
depending on how this state has been reached).
Chiral gels

Fig. 3 Eect of the spin transition on the chiro-optical properties of


nanoparticles 1. (a) CD spectra of a colloidal solution of nanoparticles
(c = 0.1 mmol L 1) recorded at dierent temperatures (298327 K) in
heating mode. Individual CD bands are detected for each spin state: 344
(negative), 317 (positive), 293 (negative), 247 (positive) and 278 (negative)
nm for the LS state and 301 (positive) and 255 (negative) nm for the
HS state. (b) Dependence of CD signals on temperature at selected
wavelengths (290 and 317 nm). The curves demonstrate a memory effect
of the chiro-optical properties associated with the hysteresis of the spin
transition (e.g., at 315 K).

coordination compound if there is a source of chirality in the


system; moreover, the chromophore itself must be situated in
an asymmetric field (in our case this is expected to be caused
by introducing a chiral anion that stimulates the formation of
a chiral lattice).
Temperature dependent CD spectra of 1 are shown in Fig. 3.
At low temperatures, two positive (317 and 247 nm) and three
negative (344, 293 and 278 nm) peaks are found for the LS
species. At high temperatures, bands characteristic for the LS
state disappear and two new bands emerge (positive at 301 nm
and negative at 255 nm) signaling the formation of the HS state
species. These CD bands in the UV region should correspond to
MLCT between iron and triazole, and to n - p* transitions in
the camphorsulfonate anion (Fig. S6, ESI).
These changes of chiro-optical properties should be driven by
two dierent factors: (a) complete reorganization of the electronic

This journal is The Royal Society of Chemistry 2015

To show the versatility of this chiral switch approach for dierent


SCO materials, a [Fe(NH2trz)3](L-CSA)2CHCl3 gel 2 was prepared
in a similar way, by using chloroform instead of acetonitrile. It is a
general tendency of complexes from the iron(II)triazole family to
form gels with some aprotic solvents (alkanes, arenes, chloroform,
etc.).38 The formation of a supramolecular system can be detected
by the increase of viscosity when the concentration of the complex
increases (Fig. 4a). Microscopically this means the formation of
a network built of highly interpenetrated complex fibres with
cavities filled by chloroform molecules35 as was confirmed for
the gel prepared here (Fig. S7 and S8, ESI).
The pink colour of the gel at room temperature is caused
by the 1A1 - 1T1 dd absorption band, which, although
spin forbidden, is still detectable in concentrated solutions
(e = 55 L mol 1 cm 1). UV-Vis spectra of this gel at different
temperatures are shown in Fig. 4b. Heating of the gel is accompanied by the disappearance of a dd transition band. The molar
extinction at 540 nm as a function of temperature is shown in
Fig. 4c. For composite 2, the SCO takes place at temperatures
similar to those observed for the nanoparticles 1, but the transition
process is slightly more gradual, with much smaller hysteresis
(T1/2 = 313 Km and 310 Kk). The loss of cooperativity should be

Fig. 4 Chiral SCO [Fe(NH2trz)3](L-CSA)2CHCl3 gel, 2. (a) Dynamic viscosity


of the SCO gel as a function of complex concentration. An abrupt increase of
the viscosity as the concentration of the complex increases demonstrates
the formation of a stable supramolecular system. (b) The UV-Vis spectra
(400800 nm) at 295327 K. The occupation of the d-orbitals in the two spin
states is inserted. (c) Temperature induced decrease of the dd transition
band at 540 nm. Photographs of the gel in the two spin states are inserted.
(d) CD spectra of the gel recorded at different temperatures (293323 K).
A CD band characteristic of the LS state is detected at 550 nm (negative).

J. Mater. Chem. C, 2015, 3, 4737--4741 | 4739

View Article Online

Published on 07 April 2015. Downloaded by Johannes Gutenberg Universitaet Mainz on 17/08/2015 15:25:06.

Paper

a direct consequence of changing the size and the morphology


of the coordination framework.
The CD behaviour of 2 corroborates its electronic absorption
in the visible region: a characteristic peak for the LS form at
540 nm (negative) disappears when the gel passes to its HS
form (Fig. 4d). The CD behaviour of gel 2 is identical in the UV
region to that observed for nanoparticles 1.
It is also important that the processing of this chiral complex
is not limited to nanoparticles and gels, but can be extended to
other functional forms: films, patterns, various composites, etc.

Conclusions
In conclusion, here we oer a pathway to monodisperse, additivefree and stable colloids of SCO nanoparticles, which can be
versatile for many SCO complexes from a wide family of Fe(II)
1,2,4-triazole compounds. This can provide a route towards dierent SCO nanoparticles with a defined temperature, abruptness and
hysteresis of spin transition. Moreover, not only the chiral anions
but also the chiral ligands can be used when constructing chiral
SCO complexes. The good stability of the presented thermoswitchable chiral nanoparticles in volatile organic solvents can permit
their future processing to thin layers, patterns and nanostructures.
A drastic change of chiro-optical properties with SCO can make
these materials attractive candidates for active elements of photonic
devices operating in the UV and visible regions. The eect of chiral
memory, especially where it appears around biologically relevant
temperatures, may also find applications in dierent experiments.

Experimental
Synthetic procedures
Iron(II) L-camphorsulfonate hexahydrate Fe(L-CSA)26H2O. The
Fe(L-CSA)26H2O salt was synthesised by a procedure similar to
that proposed by D. Coucouvanis for Fe(OTs)26H2O.39 Metallic
iron powder (100 mg, 1.79 mmol, 1.67 equiv.) was added portionwise to the solution of L-camphorsulfonic acid (500 mg, 2.16 mmol,
1 equiv.) in water (2 mL). This mixture was refluxed for 5 h.
Unreacted iron powder was filtered o and the filtrate was left to
crystallize as it cooled. Green crystals of Fe(L-CSA)26H2O were
separated by filtration and dried in air. Yield 487 mg (72%). Anal.
calcd for C20H42FeO14S2: C, 38.34; H, 6.76. Found: C, 38.42; H, 6.84.
IR (KBr): 536, 613, 1054, 1173, 1738, 2984, 2962, 3435 cm 1.
Nanoparticles of [Fe(NH2trz)3](L-CSA)2 in acetonitrile (1). Two
distinct solutions were prepared: (a) iron(II) camphorsulfonate
hexahydrate (10 mg, 0.016 mmol, 1 equiv.) in DMSO (125 mL),
then 2 mL of acetonitrile were added; (b) 4-amino-1,2,4-triazole
(10 mg, 0.119 mmol, 2.5 equiv.) in DMSO (125 mL), then 2 mL
of acetonitrile were added. The two solutions were rapidly
mixed and the mixture turned into a turbid pink colloid of
nanoparticles within a few minutes. The nanoparticles were
separated by centrifugation (20 min, 9000 rpm) and washed
with acetonitrile. They could be dried in air or redispersed in
acetonitrile depending on further experiments. Anal. calcd for
C26H42FeN12O8S2: C, 40.52; H, 5.49; N, 21.81. Found: C, 40.33;

4740 | J. Mater. Chem. C, 2015, 3, 4737--4741

Journal of Materials Chemistry C

H, 5.57; N, 21.73. IR (KBr): 509, 528, 600, 623, 1043, 1097, 1170,
1195, 1229, 1636, 1745, 2956, 3124, 3277, 3460 cm 1.
Gel of [Fe(NH2trz)3](L-CSA)2 in chloroform (2). All solvents
were thermalized at 293 K. Two distinct solutions were prepared:
(a) iron(II) camphorsulfonate hexahydrate (10 mg, 0.016 mmol,
1 equiv.) in DMSO (125 mL), then 2 mL of chloroform were
added; (b) 4-amino-1,2,4-triazole (10 mg, 0.119 mmol, 2.5 equiv.)
in DMSO (125 mL), then 2 mL of chloroform were added. The two
solutions were rapidly mixed and the mixture turned into a very
viscous pink gel within 1 h.
Physical measurements
UV-Vis measurements. UV/Vis absorbance spectra were recorded
on a Cary 50 spectrophotometer that was equipped with a
thermostated cell holder. The temperature was changed at a
rate of 2 K min 1. When following a MLCT band, the mother
solution of nanoparticles was diluted 1000 times.
CD measurements. CD spectra were recorded on a Jasco J715
CD spectrometer that was equipped with a thermostated cell
holder. The temperature was changed at a rate of 2 K min 1.
Samples were thermalized prior to every measurement for 3 min.
Data were processed using JASCO software.
Magnetic susceptibility measurements. Temperature-dependent
magnetic susceptibility measurements were carried out with a
Quantum-Design MPMS-XL-5 SQUID magnetometer equipped with
a 5 T magnet over the temperature range 2370 K with a cooling
and heating rate of 2 K min 1, and a magnetic field of 0.5 T.
Diamagnetic correction for the molecule (0.000385 emu mol 1)
was applied.
ssbauer measurements. Mo
ssbauer spectra were recorded
Mo
with a 57Co source embedded in a rhodium matrix using a
ssbauer spectrometer equipped with a heating oven.
Wissel Mo
Isomer shifts are given relative to iron metal at ambient
temperature. Simulations of the experimental data were performed with Recoil software.
Transmission electron microscopy. Particle size was determined
by using a SELMI TEM-125K transmission electron microscope.
TEM samples were prepared by placing a droplet (10 mL) of colloid
1 in acetonitrile on a carbon coated copper grid. TEM microphotographs were processed with ImageJ software.
Dynamic light scattering. DLS measurements were performed on
the colloid of nanoparticles in acetonitrile with a Malvern Zetasizer
Nano ZS and processed using standard Malvern Zetasizer software.

Acknowledgements
We acknowledge Olesia I. Kucheriv for assistance with some
synthetic work and for some useful comments on the manuscript.
We also acknowledge Wojciech Nowak for assistance with CD
measurements and Inna V. Vasylenko for TEM experiments.

Notes and references


1 B. Nowacki, H. Oh, C. Zanlorenzi, H. Jee, A. Baev, P. N. Prasad
and L. Akcelrud, Macromolecules, 2013, 46, 71587165.

This journal is The Royal Society of Chemistry 2015

View Article Online

Published on 07 April 2015. Downloaded by Johannes Gutenberg Universitaet Mainz on 17/08/2015 15:25:06.

Journal of Materials Chemistry C

2 Y. Xia, Y. Zhou and Z. Tang, Nanoscale, 2011, 3, 13741382.


3 V. K. Valev, J. J. Baumberg, C. Sibilia and T. Verbiest, Adv.
Mater., 2013, 25, 25172534.
4 S. J. Tan, M. J. Campolongo, D. Luo and W. Cheng, Nat.
Nanotechnol., 2011, 6, 268276.
5 D. Zerrouki, J. Baudry, D. Pine, P. Chaikin and J. Bibette,
Nature, 2008, 455, 380382.
6 P. Zhang, K. Mehta, S. Rehman and N. Chen, Sci. Rep., 2014,
4, 4979.
7 W. R. Browne, D. Pijper, M. M. Pollard and B. L. Feringa,
Chirality at the Nanoscale, Wiley-VCH Verlag GmbH & Co.
KGaA, 2009, pp. 349390.
8 S. Reichert and B. Breit, Org. Lett., 2007, 9, 899902.
tlich, Eur. J. Inorg. Chem., 2013, 581591.
9 P. Gu
10 M. A. Halcrow, Spin-Crossover Materials: Properties and
Applications, John Wiley & Sons, 2013.
11 K. S. Murray, H. Oshio and J. A. Real, Eur. J. Inorg. Chem.,
2013, 577580.
tlich, A. B. Gaspar and Y. Garcia, Beilstein J. Org. Chem.,
12 P. Gu
2013, 9, 342391.
13 M. Sorai, Y. Nakazawa, M. Nakano and Y. Miyazaki, Chem.
Rev., 2013, 113, PR41PR122.
14 A. Rotaru, J. Dugay, R. P. Tan, I. A. Guralskiy, L. Salmon,
r, M. Respaud and
P. Demont, J. Carrey, G. Molna
A. Bousseksou, Adv. Mater., 2013, 25, 17451749.
15 H. J. Shepherd, I. A. Guralskiy, C. M. Quintero, S. Tricard,
r and A. Bousseksou, Nat. Commun.,
L. Salmon, G. Molna
2013, 4, 2607.
16 L. Salmon, G. Molnar, D. Zitouni, C. Quintero, C. Bergaud,
J.-C. Micheau and A. Bousseksou, J. Mater. Chem., 2010, 20,
54995503.
oz, A. B.
17 M. Ohba, K. Yoneda, G. Agust, M. C. Mun
Gaspar, J. A. Real, M. Yamasaki, H. Ando, Y. Nakao,
S. Sakaki and S. Kitagawa, Angew. Chem., Int. Ed., 2009,
48, 47674771.
18 P. D. Southon, L. Liu, E. A. Fellows, D. J. Price, G. J. Halder,
tard
K. W. Chapman, B. Moubaraki, K. S. Murray, J.-F. Le
and C. J. Kepert, J. Am. Chem. Soc., 2009, 131, 1099811009.
oz-Lara, A. B. Gaspar, M. C. Mun
oz, M. Arai,
19 F. J. Mun
S. Kitagawa, M. Ohba and J. A. Real, Chem. Eur. J., 2012,
18, 80138018.
oz-Lara, M. C. Mun
oz, D. Aravena,
20 Z. Arcs-Castillo, F. J. Mun
nchez-Royo, E. Ruiz, M. Ohba,
A. B. Gaspar, J. F. Sa
R. Matsuda, S. Kitagawa and J. A. Real, Inorg. Chem., 2013,
52, 1277712783.

This journal is The Royal Society of Chemistry 2015

Paper

21 G. Agust, R. Ohtani, K. Yoneda, A. B. Gaspar, M. Ohba,


nchez-Royo, M. C. Mun
oz, S. Kitagawa and J. A. Real,
J. F. Sa
Angew. Chem., Int. Ed., 2009, 48, 89448947.
r, L. Salmon, W. Nicolazzi, F. Terki and
22 G. Molna
A. Bousseksou, J. Mater. Chem. C, 2014, 2, 13601366.
r, L. Salmon and W. Nicolazzi,
23 A. Bousseksou, G. Molna
Chem. Soc. Rev., 2011, 40, 33133335.
n-Mascaro
s, M. Monrabal-Capilla,
24 E. Coronado, J. R. Gala
n
ez, Adv. Mater., 2007,
J. Garca-Martnez and P. Pardo-Iba
19, 13591361.
25 T. Forestier, A. Kaiba, S. Pechev, D. Denux, P. Guionneau,
C. Etrillard, N. Daro, E. Freysz and J. F. Letard, Chem. Eur. J.,
2009, 15, 61226130.
26 C. Thibault, G. Molnar, L. Salmon, A. Bousseksou and
C. Vieu, Langmuir, 2010, 26, 15571560.
27 C. Faulmann, J. Chahine, I. Malfant, D. de Caro, B. Cormary
and L. Valade, Dalton Trans., 2011, 40, 24802485.
28 F. Prins, M. Monrabal-Capilla, E. A. Osorio, E. Coronado
and H. S. van der Zant, Adv. Mater., 2011, 23, 15451549.
r, I. O. Fritsky, L. Salmon and
29 I. A. Guralskiy, G. Molna
A. Bousseksou, Polyhedron, 2012, 38, 245250.
`re, A. Gloter, O. Ste
phan and
30 F. Volatron, L. Catala, E. Rivie
T. Mallah, Inorg. Chem., 2008, 47, 65846586.
ez,
31 I. Boldog, A. B. Gaspar, V. Martnez, P. Pardo-Iban
tlich and J. A. Real,
V. Ksenofontov, A. Bhattacharjee, P. Gu
Angew. Chem., Int. Ed., 2008, 47, 64336437.
32 A. Tokarev, J. Long, Y. Guari, J. Larionova, F. Quignard,
r, L. Salmon and
P. Agulhon, M. Robitzer, G. Molna
A. Bousseksou, New J. Chem., 2013, 37, 3420.
33 Y. Sunatsuki, Y. Ikuta, N. Matsumoto, H. Ohta, M. Kojima,
S. Iijima, S. Hayami, Y. Maeda, S. Kaizaki, F. Dahan and J.-P.
Tuchagues, Angew. Chem., Int. Ed., 2003, 42, 16141618.
34 W. Liu, X. Bao, L.-L. Mao, J. Tucek, R. Zboril, J.-L. Liu, F.-S. Guo,
Z.-P. Ni and M.-L. Tong, Chem. Commun., 2014, 50, 40594061.
r, I. O. Fritsky,
35 I. A. GuralSkiy, C. M. Quintero, G. Molna
L. Salmon and A. Bousseksou, Chem. Eur. J., 2012, 18,
99469954.
36 A. Grosjean, N. Daro, B. Kaumann, A. Kaiba, J.-F. Letard
and P. Guionneau, Chem. Commun., 2011, 47, 1238212384.
37 L. G. Lavrenova and O. G. Shakirova, Eur. J. Inorg. Chem.,
2013, 670682.
38 P. Grondin, O. Roubeau, M. Castro, H. Saadaoui, A. Colin
rac, Langmuir, 2010, 26, 51845195.
and R. Cle
39 D. Coucouvanis, Inorganic Syntheses, John Wiley & Sons,
Inc., New York, USA, 2002, vol. 33, pp. 75121.

J. Mater. Chem. C, 2015, 3, 4737--4741 | 4741

You might also like