ArticlePDF Available

Abstract and Figures

The autocatalytic sonochemical reaction of Fe(CO)(5) decomposition in [BuMeIm][Tf(2)N] provides iron nanoparticles in higher yields than in tetralin. Such a difference is explained by the higher decomposition of the intermediate Fe(3)(CO)(12) according to the two-sites model of the sonochemical reactions and the specific properties of the ionic liquid.
Content may be subject to copyright.
This journal is cthe Owner Societies 2011 Phys. Chem. Chem. Phys., 2011, 13, 2111–2113 2111
Cite this:
Phys. Chem. Chem. Phys
., 2011, 13, 2111–2113
Autocatalytic sonolysis of iron pentacarbonyl in room temperature ionic
liquid [BuMeIm][Tf
2
N]w
Lenaı
¨c Lartigue,
a
Rachel Pflieger,
b
Sergey I. Nikitenko,*
b
Yannick Guari,*
a
Lorenzo Stievano,
c
Moulay T. Sougrati
c
and Joulia Larionova
a
Received 1st September 2010, Accepted 8th November 2010
DOI: 10.1039/c0cp01670e
The autocatalytic sonochemical reaction of Fe(CO)
5
decomposi-
tion in [BuMeIm][Tf
2
N] provides iron nanoparticles in higher
yields than in tetralin. Such a difference is explained by the
higher decomposition of the intermediate Fe
3
(CO)
12
according
to the two-sites model of the sonochemical reactions and the
specific properties of the ionic liquid.
Synthesis of nanoparticles (NPs) under the effect of power
ultrasound is a rapidly growing area of nanochemistry since
this approach offers a versatile synthetic tool for nanostructured
materials that are often unavailable by conventional methods.
1,2
Chemical activity of ultrasound originates from acoustic
cavitation: nucleation, growth, and implosion of microbubbles
in liquids subjected to the ultrasonic waves. Collapse of cavi-
tation bubbles under acoustic resonance creates extreme local
heating inside the bubble and a thin liquid shell surrounding
the bubble. The first example of the sonochemical synthesis of
NPs was a preparation of amorphous iron during sonolysis of
Fe(CO)
5
in alkanes.
3
In the absence of stabilizers, this reaction
yields agglomerated iron NPs while an addition of stabilizers
before sonolysis (e.g. oleic acid or polyvinyl-pyrrolidone)
causes formation of colloidal iron NPs with the average size
in the range of 3–8 nm. In general, size control of iron NPs is
considered to be relatively difficult.
4
Room-temperature ionic liquids (RTILs) are known as
alternative green solvents which present unique physico-
chemical properties such as high thermal stability, negligible
vapour pressure, good ionic conductivity, large electrochemical
window, and others.
5
For these reasons, ionic liquids are
actively being explored as an environmentally benign solvent
for organic chemical reactions,
6
separations,
7
electrochemical
applications,
8
biopolymers
9
and molecular self-assemblies.
10
In the recent few years, RTILs have also been discovered as an
excellent media in the formation and stabilization of inorganic
nanosized objects
11–13
but only few works have been devoted
to the synthesis of metallic nanoparticles.
14–18
According to
Derjaguin–Landau–Verwey–Overbeek theory,
15
ionic liquids
provide an electrostatic protection in the form of a protective
shell for nanoparticles and no additional ligands or stabilizing
agents are needed. Recently, various metallic NPs have been
prepared by thermal or photochemical decomposition of metal
carbonyls M
x
(CO)
y
in imidazolium RTILs.
16,17
However, this
method often provides NPs with broad size and shape distri-
bution that is particularly true for 3d metallic NPs. Among
several examples concerning the synthesis of inorganic nano-
particles in ionic liquids, only one is devoted to the synthesis of
iron nanoparticles performed by thermal flash decomposition
up to 250 1CofFe
2
(CO)
9
in 1-methyl-3-methyl imidazolium
tetrafluoroborate.
18
Highly aggregated NPs of ca. 5 nm were
obtained and no study of their magnetic properties have been
performed. To the best of our knowledge, iron NPs have never
been obtained in RTILs by using sonochemical synthesis.
This paper describes the autocatalytic sonochemical reaction
of Fe(CO)
5
decomposition in ionic liquid [BuMeIm][Tf
2
N]
(Scheme S1, ESIw), Tf
2
N
is bis(trifluorosulfonyl)imide,
providing iron NPs with narrow size distribution. The influence
of various parameters such as the ionic liquid counter ion
nature, the Fe(CO)
5
concentration and the reaction time was
considered. Several experiments were also performed using
tetralin as a conventional hydrocarbon solvent for comparison.
In a typical experiment 10 mL of 0.5 M Fe(CO)
5
solution in
anhydrous [BuMeIm][Tf
2
N] was placed in a round bottom
Schlenk tube (1 mm wall thickness) using an argon filled inert
glove-box. Then solution was sonicated with 20 kHz ultra-
sound at an absorbed specific acoustic power P
ac
equal to
0.53 W mL
1
with a cap-horn system (Fig. S1, ESIw) under
argon flow (100 mL min
1
). A steady-state temperature inside
the reaction vessel during sonolysis was maintained at 5 1Cby
circulation of the cooled silicon oil through a cap-horn cell.
The steady-state temperature is achieved after approximately
15 minutes of ultrasonic treatment. In a few minutes after
beginning of sonolysis the reaction mixture becomes black in
both studied solvents, RTILs and tetralin, indicating an
effective sonochemical reaction. However, integral curves of
a
Institut Charles Gerhardt Montpellier, UMR 5253,
Chimie Mole
´culaire et Organisation du Solide,
Universite
´Montpellier II, Place E. Bataillon, F-34095,
Montpellier Cx5, France. E-mail: yannick.guari@univ-montp2.fr;
Fax: +33 4 67 14 38 52
b
Institut de Chimie Se
´parative de Marcoule, UMR 5257,
Centre de Marcoule, BP 17171, F-30207 Bagnols sur Ce
`ze, France.
E-mail: serguei.nikitenko@cea.fr; Fax: +33 4 67 79 76 11
c
Institut Charles Gerhardt Montpellier, UMR 5253, Agre
´gats,
Interfaces et Mate
´riaux pour l’Energie, Universite
´Montpellier II,
Place E. Bataillon, F-34095, Montpellier Cx5, France
wElectronic supplementary information (ESI) available: Experimental
details, kinetic and magnetic curves. See DOI: 10.1039/c0cp01670e
PCCP Dynamic Article Links
www.rsc.org/pccp COMMUNICATION
Downloaded by University of Montpellier on 28 January 2011
Published on 10 December 2010 on http://pubs.rsc.org | doi:10.1039/C0CP01670E
View Online
2112 Phys. Chem. Chem. Phys., 2011, 13, 2111–2113 This journal is cthe Owner Societies 2011
CO emission shown in Fig. 1 reveal the significant difference in
kinetics of CO release for these two kinds of solvent. For
tetralin solution concentration of CO in outlet gas increases
rapidly after beginning of sonolysis and reaches a steady-state
value in approximately 60 min of the process. After B100 min
of ultrasonic treatment the concentration of CO in outlet gas
decreases due to the Fe(CO)
5
consumption. By contrast, for
[BuMeIm][Tf
2
N] solution given as an example the kinetic
curve of CO emission exhibits a long induction period before
reaching a maximum after almost 4 hours of sonolysis. The
strong oscillations of CO emission during sonolysis of both
systems most probably are related to the temporal fluctuation
of bubbles coalescence which provides degassing of the liquids.
The total amount of CO released from [BuMeIm][Tf
2
N] is
almost twice (CO
RTIL
/CO
Tetralin
= 1.7) compared to that from
tetralin. Note here that the mass spectra of outlet gas also
indicate the presence of Fe and FeCO species originated
from Fe(CO)
5
evaporation. However, their concentration is
negligible compared to that of CO issued from Fe(CO)
5
sonolysis.
The sonicated solution of tetralin has a deep green colour
after the solids removal. The IR spectrum of this solution
shows the absorption bands at 2010–2050 and 1830–1870 cm
1
characteristic for C–O stretch vibrations of terminal and bridging
carbonyl groups, respectively, in a complex Fe
3
(CO)
12
.
19
This
observation is in line with the previously published data.
20,21
In
contrast to tetralin, [BuMeIm][Tf
2
N] solution presents a slight
yellow-green colour after sonolysis. Only a weak IR absorption
of Fe
3
(CO)
12
complex is observed in this case. Formation of
Fe
3
(CO)
12
complex during Fe(CO)
5
sonolysis in alkanes was
assigned to the process inside the cavitation bubbles:
3,20
4Fe(CO)
5
-))) -Fe + Fe
3
(CO)
12
+ 8CO (1)
where the symbol ‘‘-)))-’’ corresponds to the reactions under the
effect of acoustic cavitation. An increase in the bulk temperature
of the sonicated solution causes thermal decomposition of
Fe
3
(CO)
12
in solution even in the absence of ultrasound:
21
Fe
3
(CO)
12
-3Fe + 12CO (2)
Reaction (1) is known to be limited by the diffusion of
Fe(CO)
5
to the vicinity of the cavitation bubble.
21
Presumably
the diffusion coefficients of the species in [BuMeIm][Tf
2
N]
would be much lower than those in tetralin since the viscosity
of RTIL is almost 10 times higher than that of tetralin.
5
Moreover, we found that Fe
3
(CO)
12
has a low solubility in
[BuMeIm][Tf
2
N]. Therefore, the autocatalytic reaction in
RTIL can be explained by a slow reaction (1) at the first stage
of the process, accumulation of insoluble Fe
3
(CO)
12
inter-
mediate at the solution/bubble interface and its further
thermolysis in the liquid reaction zone of the cavitation
bubble according to the two-sites model of the sonochemical
reactions.
20
High solubility of Fe
3
(CO)
12
in tetralin and its low
volatility at 5 1C avoids the sonochemical decomposition of
this complex in tetralin.
About 120 (49%) and 72 mg (29%) of black magnetic solids
are removed after sonolysis of Fe(CO)
5
in [BuMeIm][Tf
2
N]
and tetralin respectively. The ratio in reaction yields of solid
products (1.7) fits very well with the ratio of released CO in
these systems. Note that decreasing the sonolysis time induces
a decrease of the nanoparticles yield (see ESIw). The proposed
reaction mechanism explains the higher reaction yield in
[BuMeIm][Tf
2
N] compared to ordinary alkanes by an addi-
tional sonochemical reaction with Fe
3
(CO)
12
.
The TEM images of the solids removed from RTIL reveal
the presence of non-aggregated uniform nanoparticles with a
mean size of 3.00(0.80) nm (Fig. 2a). Large agglomerates of
iron NPs usually observed in alkanes in the absence of
stabilizers are absent in the case of [BuMeIm][Tf
2
N]. The
X-ray diffraction pattern performed for these NPs clearly
shows the presence of bcc structure of iron (Fig. 2b).
22 57
Fe
Mo
¨ssbauer spectroscopy performed on this sample also
indicates the presence of iron nanoparticles of about 4 nm,
together with minor amounts of oxidised iron species which
might be present as oxidic iron species covering the surface of
the iron nanoparticles (Fig. S2 and Table S1, ESIw). The IR
spectra of the NPs obtained in RTIL and washed with
anhydrous THF clearly indicate the bands at 1572 cm
1
(nCQC), 1465 cm
1
(d
S
CH
3
), 1430 cm
1
(d
S
CH
3
(Me)),
1198 cm
1
(nN-Bu) and 1184 cm
1
(nN-Me), 1056 (nC–C) of
BuMeIm
+
cation from RTIL. This observation confirms that
RTILs are effective stabilizers for NPs. Moreover, these NPs
may be re-dispersed in RTILs (Fig. S3a, ESIw).
Using RTILs with [BF
4
]or[PF
6
] as counter anions instead
of [Tf
2
N
] also allows the synthesis of iron NPs with similar
sizes (Fig. S3b, ESIw) but with lower yields, which can be
explained by the lower solubility of Fe(CO)
5
in these RTILs.
Using higher Fe(CO)
5
concentration in [BuMeIm][Tf
2
N] leads
to the formation of iron cauliflowers (Fig. S3c, ESIw).
The magnetic properties of the NPs obtained in
[BuMeIm][Tf
2
N] studied by using static (dc) and dynamic
(ac) modes confirm the presence of iron NPs. The Zero
Fig. 1 (a) Integrated kinetic curves for CO emission during sonolysis
of 0.5 M Fe(CO)
5
in tetralin (grey) and [BuMeIm][Tf
2
N] (black)
(20 kHz, Pac = 0.53 W mL
1
,T=51C, Ar) and (b) the corres-
ponding raw data.
Fig. 2 (a) TEM image of iron NPs obtained in [BuMeIm][Tf
2
N]. The
TEM samples were prepared by suspending the NPs in ethanol and
deposition of a drop on a copper grid. (b) X-Ray diffraction of the NPs
obtained in [BuMeIm][Tf
2
N].
}
shows the lines originated from the
confined dome-like holder equipped with a stainless knife edge beam.
Downloaded by University of Montpellier on 28 January 2011
Published on 10 December 2010 on http://pubs.rsc.org | doi:10.1039/C0CP01670E
View Online
This journal is cthe Owner Societies 2011 Phys. Chem. Chem. Phys., 2011, 13, 2111–2113 2113
Field-Cooled (ZFC) magnetization curve shows a peak at
T
max
= 184 K, which is typical for iron NPs of this size
(Fig. 3 and Fig. S4, ESIw).
23
The FC curve increases, reaches a
maximum value at 125 K then decreases suggesting the presence
of superspin-glass behaviour characteristic of interacting NPs.
24
The field dependence of the magnetization measured at 2.5 K
shows the presence of a hysteresis effect with a value of the
coercive field of 535 Oe and the value of the saturation magneti-
zation of 142 emu g
1
(inset of Fig. 3). As the temperature
increases, the coercive field decreases and at 300 K the hysteresis
loop is closed (Fig. S5, ESIw).
In order to determine the nature of the magnetic behaviour of
these nanoparticles, the dynamic ac magnetic measurements were
performed. The temperature dependence of the in-phase (w0)and
out-of-phase (w00) components of the ac susceptibility measured at
different frequencies shows a frequency dependent behaviour
(Fig. S6, ESIw). The thermal dependence of the relaxation time
can be described by a critical scaling law, t=t
0
[T
g
/(TT
g
)]
zv
(T
g
is the glass temperature, and zv is a critical exponent)
25
with
the satisfactory fit parameters: T
g
= 155 K, t
0
=1.2910
6
s
and zn= 11.5 (inset of Fig. S6, ESIw), indicating the presence of
superspin-glass behavior induced by strong magnetostatic inter-
actions between the nanoparticles.
26
The field dependence of
the spin-glass transition temperature, T
g
, taken from the field
dependence of the ac susceptibility (Fig. S7, ESIw)followsthede
Almeida–Thouless line,
27
HN[1 (T
max
/T
g
)]
3/2
(Fig. S8, ESIw)
confirming the existence of a spin-glass phase.
28
In summary, we describe the sonochemical autocatalytic
reaction of Fe(CO)
5
in various RTILs. This reaction produces
with a relatively high yield considering a sonolysis time of
5 h in [BuMeIm][Tf
2
N] non-agglomerated ultra small iron
nanoparticles of ca. 3.0 nm with a narrow size distribution.
The study of the magnetic properties of these nanoparticles by
using static and dynamic analyses reveals a superspin-glass like
behaviour of these iron nanoparticles induced by the presence
of interparticles interactions.
Notes and references
1 J. H. Bang and K. S. Suslick, Adv. Mater., 2010, 22, 1039.
2 A. Gedanken, Ultrason. Sonochem., 2004, 11, 47.
3 K. S. Suslick, M. Fang and T. Hyeon, J. Am. Chem. Soc., 1996,
118, 11960.
4 D. L. Huber, Small, 2005, 1, 482.
5(a)T.Welton,Chem. Rev., 1999, 99, 2071; (b) P. Wasserscheid and
W. Keim, Angew. Chem., Int. Ed., 2000, 39,3772;(c)L.A.Blanchard,
D. Hancu, E. J. Beckman and J. F. Brennecke, Nature, 1999, 399,28.
6(a) P. Wasserscheid and T. Welton, Ionic Liquids in Synthesis,
Wiley-VCH, Weinheim, 2003; (b) R. Scheldon, Chem. Commun.,
2001, 2399.
7 J. G. Huddleston, H. D. Willauer, R. P. Swatloski, A. E. Visser
and R. D. Rogers, Chem. Commun., 1998, 1765.
8(a) A. B. McEwen, S. F. McDevitt and V. R. Koch, J. Electrochem.
Soc., 1997, 144, L84; (b) E. V. Dickinson, M. E. Williams,
S. M. Hendrickson, H. Masui and R. W. Murray, J. Am. Chem.
Soc., 1999, 121, 613.
9(a) N. Kimizuka and T. Nakashima, Langmuir, 2001, 17, 6759;
(b) R. P. Swatloski, S. K. Spear, J. D. Holbrey and R. D. Rogers,
J. Am. Chem. Soc., 2002, 124, 4974.
10 T. Nakashima and N. Kimizuka, Chem. Lett., 2002, 1018.
11 G. Clavel, J. Larionova, Y. Guari and Ch. Gue
´rin, Chem.–Eur. J.,
2006, 12, 3798.
12 (a) C. W. Scheeren, G. Machado, J. Dupont, P. F. P. Fichtner and
S. R. Texeira, Inorg. Chem., 2003, 42, 4738; (b) J. Dupont and
J. D. Scholten, Chem. Soc. Rev., 2010, 39, 1780; (c) Y. Wang,
S. Maksimuk, R. Shen and H. Yang, Green Chem., 2007, 9, 1051.
13 (a) J. Huang, T. Jiang, B. Han, H. Gao, Y. Chang, G. Zhao and
W. Wu, Chem. Commun., 2003, 1654; (b) Y.-J. Zhu, W.-W. Wang,
R.-J. Qi and X.-L. Hu, Angew. Chem., Int. Ed., 2004, 43, 1410;
(c) T. Nakashima and N. Kimizuka, J. Am. Chem. Soc., 2003, 125,
6386.
14 M. Scariot, D. O. Silva, J. D. Scholten, G. Machado,
S. R. Teixeira, M. A. Novak, G. Ebeling and J. Dupont, Angew.
Chem., Int. Ed., 2008, 47, 9075.
15 E. J. W. Verwey and J. T. G. Overbeek, Theory of the Stability of
Lyophobic Colloids, Dover Publications, Mineola, NY, 2nd edn,
1999.
16 E. Redel, R. Thomann and C. Janiak, Chem. Commun., 2008,
1789.
17 D. O. Silva, J. D. Scholten, M. A. Gelesky, S. R. Teixeira, A. C. B.
Dos Santos, E. F. Souza-Aguiar and J. Dupont, ChemSusChem,
2008, 1, 291.
18 J. Kra
¨mer, E. Redel, R. Thomann and C. Janiak, Organometallics,
2008, 27, 1976.
19 K. Nakanishi and H. Solomon, Infrared Absorption Spectroscopy,
Holden-Day, Inc., San Francisco, 1982.
20 K. S. Suslick, S. B. Choe, A. A. Cichowlas and M. W. Grinstaff,
Nature, 1991, 353, 414.
21 S. I. Nikitenko, Yu. Koltypin, I. Felner, I. Yeshurun, A. I. Shames,
J. Z. Jiang, V. Markovich, G. Gorodetsky and A. Gedanken,
J. Phys. Chem. B, 2004, 108, 7620.
22 J. Carvella, E. Ayietaa, M. Johnsona and R. Cheng, Mater. Lett.,
2009, 63, 715.
23 (a) M. Yoona, Y. M. Kima, Y. Kima, V. Volkova, H. J. Songa,
Y. J. Park, S. L. Vasilyakc and I.-W. Park, J. Magn. Magn. Mater.,
2003, 265, 357; (b) G. Kataby, Y. Koltypin, A. Ulman, I. Felner
and A. Gedanken, Appl. Surf. Sci., 2002, 201, 191.
24 D. Parker, V. Dupuis, F. Ladieu, J.-P. Bouchaud, E. Dubois,
R. Perzynski and E. Vincent, Phys. Rev. B: Condens. Matter
Mater. Phys., 2008, 77, 104428.
25 J. A. Mydosh, Spin Glasses, Taylor and Francis, Washington, DC,
1993.
26 C. Djurberg, P. Svedlindh, P. Nordblad, M. F. Hansen, F. Bodker
and S. Morup, Phys. Rev. Lett., 1997, 79, 5154.
27 J. R. L. Almeida and D. J. Thouless, J. Phys. A: Math. Gen., 1978,
11, 983.
28 B. Martinez, X. Obradors, Ll. Balcells, A. Rouanet and C. Monty,
Phys. Rev. Lett., 1998, 80, 181.
Fig. 3 Field Cooled (FC) (grey)/Zero Field Cooled (ZFC) (black)
magnetization curves performed under an applied magnetic field of
1000 Oe. Inset: hysteresis loops of iron NPs performed at 2.5, 50, 150
and 300 K.
Downloaded by University of Montpellier on 28 January 2011
Published on 10 December 2010 on http://pubs.rsc.org | doi:10.1039/C0CP01670E
View Online
... Surprisingly, the thermal decomposition of iron carbonyl complexes in pure ILs has been little investigated and to our knowledge, only two examples can be mentioned: first, the thermal flash decomposition of Fe 2 (CO) 9 up to 250°C in 1-methyl-3-butylimidazolium tetrafluoroborate which leads to the formation of highly aggregated iron NPs of ca. 5 nm (Krämer et al. 2008). Second, the sonochemical decomposition of Fe(CO) 5 in 1-methyl-3-butyl imidazolium salts leading to the formation of non-aggregated spherical iron NPs with a narrow size distribution centered at 2 nm (Lartigue et al. 2011). In both cases, the formation of iron carbides was not reported. ...
Article
The thermal decomposition of Fex (CO)y precursors for the synthesis of nanoparticles of iron carbides and their superstructures with sizes ranging from 2.8 to 15.1 nm is developed using imidazolium-based ionic liquids as solvents, stabilizers, and carbon source. A study of the influence of some synthesis parameters such as the heating temperature, nature, and concentration of the iron carbonyl precursor and chain length of the N-alkyl substituent on the imidazolium ring on the size and organization of the iron carbide nanoparticles is presented. These iron carbides nano-objects were characterized by infra-red spectroscopy, transmission electronic microscopy, powder X-ray diffraction, Mossbauer spectroscopy, and magnetic analyses.
Article
Sb2S3 nanorods were successfully synthesized by the ionic liquid assisted sonochemical method (ILASM). The starting reagents were Sb2Cl3, Thioacetamide, absolute ethanol (ETA) and the new ionic liquid used was isobutyl-3 hexyl-phosphonium tetrafluoroborate ([iBH3P][BF4]). The synthesized materials were subjected to 120°C annealing temperature under controlled vacuum conditions. X ray powder diffraction analysis showed that the ionic liquid played a key role in the cristallinity of Sb2S3. Scanning electron microscopy displayed single-crystalline Sb2S3 nanorods that could be prepared in the presence of [iBH3P][BF4]. Energy-dispersive X-ray spectroscopy confirmed the formation of Sb2S3.The optical properties (band gap) were similar to that reported for bulk Sb2S3.
Article
Nanostructured transition metal (Fe, Co and Ni) carbides and nitrides have attracted much attention due to their active properties in various areas such as nanomagnetism, biomedicine, catalysts for electrochemistry, and environmental architectures. Generally, the intrinsically harsh characteristics of carbides and nitrides make their synthetic protocols difficult, especially at the nanoscale, but usually with improved performances for further specific applications. In this review article, we summarized the iron, cobalt and nickel carbides and nitrides nanostructures focused on the interactions between electronic, magnetic structures and the related crystallographic structures, and the traditional or de novosynthetic strategies, and their various applications from biomedical to environmental architectures.
Article
SnS nanoparticles have been successfully synthesized by the ionic liquid-assisted sonochemical method (ILASM). The starting reagents were anhydrous SnCl2, thioacetamide, dissolved in ethanol and ionic liquid (IL) 1-butyl-3-methylimidazolium tetrafluoroborate (BMImBF(4)) mixtures. Our experiments showed that IL plays an important role in the morphology of SnS. A 1: 1 ethanol: IL mixture was found to yield the more interesting features. The lower concentration of Sn (II) in solution favored the presence of nanoplatelets. An increase in ultrasonic time favored crystalline degree and size as well. Also, the effect of additives as 3-mercaptopropionic acid, diethanolamine, ethylene glycol, and trioctyl phosphine oxide is reported. X-ray diffraction (XRD) and ultraviolet-visible diffuse reflectance spectroscopy (UV-Vis-DRS) were used to characterize the obtained products.
Article
Doubtless ionic liquids (ILs), particularly those based on the 1,3-dialkyl imidazolium cation, provide a flexible liquid platform to prepare, soluble and stable transition metal nanoparticles (TMNPs). ILs can act as a “solvent”, stabiliser, ligand and support for TMNPs. Soluble and stable TMNPs for specific applications can be easily prepared in ILs using a bottom-up or top-down approach. The stability of TMNPs in non-functionalised ILs is mainly related to the surface electronic stabilisation provided by protective layers of discrete supramolecular imidazolium aggregates, non-polar imidazolium alkyl side chains and NHC carbene species as well as surface hydrogen species, together with an oxide layer when present on the metal surface. The IL provides a template-like effect and does not exist as a pure double layer, rather, the IL interacts directly with the TMNP surface through both cationic and anionic species, and the non-polar groups are preferentially directed away from the surface, forming a protective layer at the interface of at least one layer thick. The main aspects involving the stabilisation, in particular the interface of TMNP surface with the ionic liquids and other species present in the media, will be presented and discussed in light of the recent experimental and theoretical results reported.
Article
Full-text available
Ionic liquids (ILs)-stabilized iron oxide (Fe(2)O(3)) nanoparticles were synthesized by the ultrasonic decomposition of iron carbonyl precursors in [EMIm][BF(4)] without any stabilizing or capping agents. The Fe(2)O(3) nanoparticles were isolated and characterized by X-ray powder diffraction, transmission electron microscopy and susceptibility measurements. The physicochemical properties of ILs containing magnetic Fe(2)O(3) nanoparticles (denoted as Fe(2)O(3)@[EMIm][BF(4)]), including surface properties, density, viscosity and stability, were investigated in detail and compared with that of [EMIm][BF(4)]. The Fe(2)O(3)@[EMIm][BF(4)] can be directly used as magnetic ionic liquid marble by coating with hydrophobic and unreactive polytetrafluoroethylene (PTFE), for which the effective surface tension was determined by the puddle height method. The resulting magnetic ionic liquid marble can be transported under external magnetic actuation, without detachment of magnetic particles from the marble surface that is usually observed in water marble.
Article
Full-text available
Phenanthroline (Phen) ligand-protected palladium nanoparticles in ionic liquid (IL) 1-n-butyl-3-methylimidazolium hexafluorophosphate ([BMIM][PF6]) are very active and selective for the hydrogenation of olefins, and the nanoparticles/IL system can be reused many times without reducing the activity.
Article
Full-text available
In this paper we investigate the superspin glass behavior of a concentrated assembly of interacting maghemite nanoparticles and compare it to that of canonical atomic spin glass systems. ac versus temperature and frequency measurements show evidence of a superspin glass transition taking place at low temperature. In order to fully characterize the superspin glass phase, the aging behavior of both the thermo-remanent magnetization (TRM) and ac susceptibility has been investigated. It is shown that the scaling laws obeyed by superspin glasses and atomic spin glasses are essentially the same, after subtraction of a superparamagnetic contribution from the superspin glass response functions. Finally, we discuss a possible origin of this superparamagnetic contribution in terms of dilute spin glass models.
Article
Ionic liquids are salts that are liquid at low temperature (<100 °C) which represent a new class of solvents with nonmolecular, ionic character. Even though the first representative has been known since 1914, ionic liquids have only been investigated as solvents for transition metal catalysis in the past ten years. Publications to date show that replacing an organic solvent by an ionic liquid can lead to remarkable improvements in well‐known processes. Ionic liquids form biphasic systems with many organic product mixtures. This gives rise to the possibility of a multiphase reaction procedure with easy isolation and recovery of homogeneous catalysts. In addition, ionic liquids have practically no vapor pressure which facilitates product separation by distillation. There are also indications that switching from a normal organic solvent to an ionic liquid can lead to novel and unusual chemical reactivity. This opens up a wide field for future investigations into this new class of solvents in catalytic applications.
Article
Ionic liquids are salts that are liquid at low temperature (<100°C) which represent a new class of solvents with nonmolecular, ionic character. Even though the first representative has been known since 1914, ionic liquids have only been investigated as solvents for transition metal catalysis in the past ten years. Publications to date show that replacing an organic solvent by an ionic liquid can lead to remarkable improvements in well-known processes. Ionic liquids form biphasic systems with many organic product mixtures. This gives rise to the possibility of a multiphase reaction procedure with easy isolation and recovery of homogeneous catalysts. In addition, ionic liquids have practically no vapor pressure which facilitates product separation by distillation. There are also indications that switching from a normal organic solvent to an ionic liquid can lead to novel and unusual chemical reactivity. This opens up a wide field for future investigations into this new class of solvents in catalytic applications.
Article
Interesting ionic materials can be transformed into room temperature molten salts by combining them with polyether-tailed counterions such as polyether-tailed 2-sulfobenzoate (MePEG-BzSO{sub 3}{sup {minus}}) and polyethertailed triethylammonium (MePEG-Et{sub 3}N{sup +}). Melts containing ruthenium hexamine, metal trisbipyridines, metal trisphenanthrolines, and ionic forms of aluminum quinolate, anthraquinone, phthalocyanine, and porphyrins are described. These melts exhibit ionic conductivities in the 7 x 10{sup {minus}5} to 7 x 10{sup {minus}10} {Omega}{sup {minus}1} cm{sup {minus}1} range, which permit microelectrode voltammetry in the undiluted materials, examples of which are presented.
Article
The partitioning of simple, substituted-benzene derivatives between water and the room temperature ionic liquid, butylmethylimidazolium hexafluorophosphate, is based on the solutes’ charged state or relative hydrophobicity; room temperature ionic liquids thus may be suitable candidates for replacement of volatile organic solvents in liquid–liquid extraction processes.
Article
Dialkyldimethylammonium amphiphiles form stable bilayer membranes in the ether-containing ionic liquids. Dark field optical microscopy indicates the formation of vesicles above the phase transition temperature.
Article
This paper reports the synthesis of iron oxide nanoparticles using a freshly-made or recycled 1-butyl-3-methylimidazolium bis(triflylmethylsulfonyl)imide ([BMIM][Tf2N]) ionic liquid (IL). Iron pentacarbonyl (Fe(CO)5), which dissolves in [BMIM][Tf2N], thermally decomposed and subsequently oxidized to form iron oxide nanoparticles. These nanoparticles separated out automatically from the imidazolium-based ionic liquid mixtures. Multiple additional runs were tested in making iron oxide nanoparticles using recycled ionic liquid. The iron oxide nanoparticles made were characterized with transmission electron microscopy (TEM), high resolution TEM (HR-TEM) and powder X-ray diffraction (PXRD). The structure and thermal stability of the IL was examined using Fourier transform infrared (FT-IR) spectroscopy and thermal gravimetric analysis (TGA). We found that iron oxide nanoparticles with a narrow size distribution could be obtained. The [BMIM][Tf2N] ionic liquid showed no degradation based on the TGA and FT-IR study. The solvent-recyclable process of making size-controlled nanoparticles should have a broad impact on the application of imidazolium-based ionic liquids in the synthesis of nanomaterials.