Sunteți pe pagina 1din 5

Optical Materials 34 (2011) 336340

Contents lists available at ScienceDirect

Optical Materials
journal homepage: www.elsevier.com/locate/optmat

Highly doped alkaline earth nanouorides synthesized from ionic liquids


C. Lorbeer a, J. Cybinska a,b, E. Zych b, A.-V. Mudring a,
a
b

Ruhr-Universitt Bochum, Anorganische Chemie I Festkrperchemie und Materialien, D-44780 Bochum, Germany
University of Wrocaw, Faculty of Chemistry, F. Joliot-Curie 14, PL-50383 Wrocaw, Poland

a r t i c l e

i n f o

Article history:
Received 17 February 2011
Accepted 11 April 2011
Available online 20 May 2011
Keywords:
Alkaline earth uorides
Quantum cutting
Ionic liquids
Morphology
Nanoparticles

a b s t r a c t
Bulk alkaline earth uorides co-doped with optically active ions are prominent materials for luminescent
applications. However, for phosphor materials the changeover to the nanoscale is a tightrope walk
between achieving desirable features of small particles such as reduced light scattering and unwanted
drawbacks such as a high surface defect concentration which is likely to result in quenching of luminescence. A new preparation route via ionic liquids allows obtaining pure and oxygen-free alkaline earth uorides co-doped with Eu3+ and Gd3+ on the nanoscale with excellent quantum cutting abilities.
2011 Elsevier B.V. All rights reserved.

1. Introduction
Nanotechnology is currently one of the most popular research
topics because of device miniaturization and the search for new
materials with exceptional properties and property combinations.
Many chemical and physico-chemical properties of a material
change from the bulk to the nanosize regime. This applies also to
optical properties and nanotechnology has also become important
in the eld of phosphor chemistry [1]. Although nanoparticles are
dened as having a size below 100 nm, really amazing properties
occur at particle sizes below 10 nm because the fraction of surface
atoms is signicant. The different behavior of surface atoms is
caused by the reduced number of neighbor atoms compared to
the saturated coordination environment of a volume atom. Furthermore, stabilities of polymorphs depend on the particle size and may
be different from the bulk [2]. Particle size may also affect physicooptical properties such as emission lifetime, quantum efciency and
concentration quenching [3]. In addition, the symmetry and crystal
eld as well as the refractive index can deviate from the bulk.
Taking a phosphor material to the nanoscale is a borderline
walk between achieving desirable features such as reduced light
scattering and unwanted troubles arising from the large surface
area. Although uoride materials are extremely interesting for
luminescent applications because of their high chemical stability,
their wide band gap, the possible high dopant ion level, the good
mechanical properties and comparatively high moisture resistance,

Corresponding author.
E-mail address: anja.mudring@rub.de (A.-V. Mudring).
0925-3467/$ - see front matter 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.optmat.2011.04.019

nanocrystalline uoride materials are commonly not used and often not explored because their favorable properties are considerably reduced by surface species such as OH or NH or oxide
contaminations introduced during the synthesis. These make efcient multiphonon relaxation possible and one of the major advantages of uoride materials, the low phonon energies, is lost. Indeed,
the synthesis of pure, quenching species-free nanouorides is
rather difcult. Recently, we developed a microwave synthesis of
nanocrystalline lanthanide uorides using ionic liquids as the solvent, reactant and in situ stabilizer, all in one [4,5]. The obtained
nanomaterials with a particle size below 10 nm had excellent
photo-physical properties and no surface quenching species could
be detected. However, lanthanide uorides are problematic with
respect to their pyro-hydrolytic and hydrolytic stability [6]. To
simultaneously improve the stability beyond lanthanide uorides
and to check generalization of our synthesis protocol, different
host materials have been considered. Alkaline earth uorides prot
from a substantially higher pyro-hydrolytic and hydrolytic stability. Besides, concentration quenching of optically active lanthanide
ions usually occurs at higher concentrations in smaller particles
and in alkaline earth uorides co-doped with rare earth ions, solid
solutions can be produced over a wide range of concentrations
[1,7,8]. Despite this, only few reports on nanocrystalline alkaline
earth uoride particles have appeared until now [912].
Alkaline earth uorides MF2 (M = Ba, Sr, Ca), crystallizing in the
cubic CaF2 type of structure [13], can be doped to a high degree
with trivalent lanthanide uorides. The lanthanide ions occupy
the M2+ sites of the host material. The substitution of a divalent
metal cation by a trivalent ion requires charge compensation
which is achieved by the introduction of additional uoride anions

C. Lorbeer et al. / Optical Materials 34 (2011) 336340

in the lattice. However, the charge compensating uoride ion can


modify the coordination environment, hence, the site symmetry
of the metal site ((4a) in Fm-3 m, Oh symmetry) [14]. X-ray structure investigations on bulk material indicate that defect clusters
form, i.e. areas form, where trivalent rare earth cations and charge
compensating extra-uoride anions aggregate. X-ray and neutron
diffraction studies have been used to investigate the defect structure of bulk materials [15]. For example, for Ba1xLnxF2+x (Ln = trivalent rare earth) the formation of [Ln6F36] groups, where the Ln3+
cations form an octahedron, was found. Depending on the alkaline
earth ion M2+ of the host and the doping Ln3+, different local symmetries for Ln3+ can be determined spectroscopically [16]. Albeit,
well investigated, the reports in the literature are contradictory.
It has been claimed that in BaF2 the compensating interstitial F 
i
ion (4b) at (, , ) is located in the nearest neighbor site (NNN
3+
manner, the Goldschmidt position) with respect to the Eu ion,
thus providing a trigonal (C3v) symmetry environment [17]. A
tetragonal (C4v) symmetry arises when the nearest neighbor interstitial F 
i ions (NN manner) are located in the h1 1 0i direction from
the Ln3+ ion. This prevails in CaF2 [18]. For SrF2 it was observed that
the metal site symmetry varies dependent on the doped lanthanide
ion. For Ce3+ to Ho3+, the optically active ion forms a C4v symmetry
center similar to that formed in CaF2. For Ln3+ ions beyond holmium, the charge compensating interstitial F 
i ion is located along
the h1 1 1i direction from the Ln3+ ion resulting in a C3v symmetry
[16]. For heavily doped MF2:Ln3+, the studies on charge compensation and the site symmetries of the optically active ion are scarce
[18]. It seems that, with increasing Ln3+ concentration, the Ln3+
F
i dipoles couple to dimers, trimers and higher aggregates. The degree of defect clustering seems to depend on the M2+/Ln3+ radii ratio. The cluster type of (M, Ln)4F26 is dominant with a r(M2+)/
r(Ln3+) ratio below 0.95, whereas at higher ratios the formation
of (M, Ln)6F36 and M8Ln6F6869 clusters prevail.
However, to the best of our knowledge, no information about
charge compensation in nanoscale alkaline earth uoride exists up
to now. Also the mechanism of charge compensation in heavily
doped alkaline earth uorides (<5%) is also only faintly studied,
probably due to the lack of application for such materials. Typically,
the smaller the particles and the higher the doping level of the optically active ion becomes the more the concentration quenching
behavior of nanoparticles deviates from that of the bulk material.
In this study, we rst synthesized Eu3+ and Gd3+ co-doped BaF2 nanoparticles with different gadolinium concentrations to analyze the
inuence of composition on materials properties such as size, crystal
phase and luminescence. The optimum Ba2+:Ln3+ was then used for
the similar synthesis of Eu3+ and Gd3+ co-doped CaF2 and SrF2.

337

hydrate, dissolved in a small amount of ethylene glycole, was


mixed with an excess of [C4mim][BF4], in a 10 ml glass vessel
equipped with a Teon septum and heated in 5 min to 120 C
with a single mode microwave operating at 2455 MHz (CEM Discover, Kamp-Lintfort, D). The reaction mixture was kept at this
temperature under continuous high stirring for additional 10 min
before cooling to room temperature by pressurized air. The obtained colloidal solutions were centrifuged and the reaction product washed several times with an ethanol/dichloromethane
mixture. Finally, the colorless powder was dried at 70 C.
2.2. Characterization
Powder X-ray diffraction measurements were carried out on a
G670 diffractometer with an image plate detector (Huber, Rimsting, D) operating with Mo Ka radiation. Calculation of strain
and particle size of the obtained material was carried out with
WinXPow 1.07 (Stoe, Darmstadt, Germany). Rietveld renements
were carried out with FULLPROF [20] using the graphical interface
WinPLOTR [21]. As starting congurations cubic [22] and tetragonal [23] structures of similar compounds were used. TEM (transmission electron microscopy) micrographs were obtained on a
Tecnai G2 20 X-Twin transmission electron microscope (Fei, Hillsboro, USA). Energy-dispersive X-ray microanalysis was carried
using an EDAX detector (Fei, Hillsboro, USA). An acceleration voltage of 200 kV was applied. For sample preparation, a suspension of
the sample in ethanol was dropped on a copper or gold grid coated
with carbon and then dried in air. The particle size distribution was
obtained by measuring the size of the synthesized particles on different overview TEM images with the program ES Vision (Fei, Hillsboro, USA) and the averaged particle size and the standard
deviation were calculated from the data. Luminescence measurements were carried out on a Fluorolog FL 3-22 spectrometer (Horiba JobinYvon, Unterhachingen, D) equipped with a continuous
xenon lamp for steady state measurements and a pulsed xenon
lamp for determining decay curves. Double gratings for the excitation and emission spectrometer are applied as monochromators.
The signal is detected by a photomultiplier. For measurement,
powdered samples were lled in silica tubes and carefully positioned in the incoming beam in the sample chamber. VUV spectroscopic measurements with synchrotron radiation were undertaken
on beamline I (SUPERLUMI station) at the Hamburger Synchrotronstrahlungslabor (HASYLAB), Deutsches Elektronensynchrotron
(DESY), Germany [24].
3. Results and discussion

2. Experimental
Synthesis and sample handling were carried out using standard
Schlenk and Argon-glove box techniques. The starting chemicals,
1-methylimidazole (99%, Sigma Aldrich), chlorobutane (99%, Acros), acetonitrile (99.5%, J.T. Baker), sodium tetrauoroborate
(98%, Aldrich), dichloromethane (99.9%, Fisher Scientic), barium
acetate (99%, ABCR), strontium acetate (99%, ABCR), calcium acetate hydrate (99%, ABCR), europium (III) acetate hydrate (99.9%,
ABCR), gadolinium (III) acetate tetrahydrate (99.9%, ABCR), ethylene glycol (99%, J.T. Baker) and ethanol (p.a., Riedel de Hen) were
used as received. The ionic liquid 1-butyl-3-methylimidazolium
tetrauoroborate, [C4mim][BF4], was synthesized according to a
common literature procedure [19].
2.1. Nanoparticle synthesis
In a typical reaction an appropriate amount of the alkaline earth
acetates, gadolinium acetate tetrahydrate and europium acetate

X-ray powder diffraction pattern (PXRD) conrms that in all


cases phase pure material was obtained (Fig. 1). BaF2 co-doped
with 1% Eu3+ and 515% Gd3+ as well as CaF2 and SrF2 co-doped
with 1% Eu3+ and 15% Gd3+ could be rened in the cubic CaF2 type
of structure. For higher Ln3+ concentrations a tetragonal distortion
of the CaF2 structure was observed and the compounds could be
rened in the space group I4/mmm with the lattice constants given
in Table 1. As expected, the molar volume of GdEu co-doped BaF2
decreases with increasing Ln3+ content. Unfortunately, for the
nanosized samples the free positional parameters of the additional
uoride ions could not be rened due to strong broadening and
extensive overlap of the reections. A deviation from the ideal cubic lattice with increasing dopant concentration is expected as
more and more divalent metal cations are exchanged by trivalent
cations and in consequence, more and more extra-uoride anions
have to be incorporated in the crystal lattice. We calculated the
contributions of lattice strain on the line broadening and found
that with increasing the gadolinium concentration from 5% to

338

C. Lorbeer et al. / Optical Materials 34 (2011) 336340

Fig. 1. Powder diffraction patterns of BaF2 co-doped with Eu3+ and different amounts of Gd3+ (left) and corresponding electron diffraction images (right).

Table 1
Lattice constants and molar volume for BaF2 co-doped with Eu 1% and Gd 525% as
obtained from Rietveld renements.
BaF2, Eu 1%, Gd (%)

a, c ()

Vm/106 (3)

5
10
15
20

6.1598(4)
6.1488(3)
6.1472(5)
4.3054(3)
6.2188(7)
4.2462(3)
6.1900(6)

58.43
58.11
58.07
57.64

25

55.81

25% a doubling of the strain in the obtained particles occurs. The


electron diffraction patterns (insert, Fig. 1) support this. The pattern of the 5% Gd3+ co-doped sample shows single crystalline features but with increasing gadolinium concentration the features of
the diffraction patterns become more and more polycrystalline.
Concurrently, the crystallite size determined with means of the
Scherrer equation from the PXRD data gets smaller from 7.9 nm
in BaF2:Eu 1%, Gd 5% down to 4.6 nm in case of the BaF2:Eu 1%,
Gd 25%.
To conrm this and to obtain information about the morphology of the synthesized particles, transmission electron microscopy
(TEM) measurements have been carried out (for representative
TEM micrographs see Fig. 2). The change of crystallite size with
the dopant concentration obtained from particle size analysis of
the TEM measurements shows the same trend as determined from
PXRD. The size of the single particles starts from 10 nm and decreases to 7.5 nm. However, as more than one crystallite domain
is present in most cases in one particle, the size determined from
TEM is slightly higher when compared to the PXRD data. In addition the TEM images reveal that with increasing gadolinium concentration, a stronger aggregation of the single particles occurs.
In case of 5% Gd3+ co-doping, we obtain clustering of the particles
to plate-like aggregates (tilting of the TEM holder reveals that the
mainly observed crosses obtain a more and more oval form). The
agglomeration proceeds with increasing Gd3+ concentration creating more and more three dimensional aggregates, nally leading to
spherical clusters of 200250 nm diameter as observed in the sample co-doped with 25% of Gd3+.
In case of SrF2 and CaF2 co-doped with 1% Eu3+ and 15% Gd3+,
the tendency to aggregate is less pronounced, but the particles
themselves are larger with an average size of 14 nm for calcium
uoride and 16 nm for strontium uoride, respectively (Fig. 2).

All samples under investigation show efcient orangered


emission even observed with the naked eye. The room temperature
emission spectra of the europium and gadolinium co-doped BaF2
7
directly excited via the 5L6
F0,1 transition of Eu3+ at kex = 393 nm
are presented in Fig. 3. The BaF2:Eu, Gd nanoparticles show a number of narrow lines corresponding to the 4f4f (5D0 ? 7FJ) transitions. In addition to the 5D0 emission, lines arising from 5DJ
emitting levels (J = 1 and 2) were also observed. With Gd3+ concentrations varying from 10% to 25%, similar emission spectra are obtained. However, with increasing gadolinium concentration the
relative intensity of emission from 5D1 level with respect to emission from 5D0 is enhanced as a result of the larger dilution of the
Eu3+ ion among the Gd3+ ions. This points to a clustering of the lanthanide ions in the BaF2 lattice, as the relative intensity of emission
from 5D1 level should not be affected if the Ln3+ ions would be randomly distributed in the BaF2 lattice. In the sample co-doped with
1% Eu3+ and 5% Gd3+, a slight change in the emission spectrum can
be observed in comparison to the heavier doped materials: While
the main character of the spectrum remains the same (intensity ratio of the 5D0 ? 7FJ transitions), the relative intensity of the Stark
components (particularly in 5D0 ? 7F2 transition) change noticeably. As it was previously discussed, a different degree of lanthanide doping, hence, defect concentration leads to different
structural changes, thus resulting in different and also additional
site symmetries for the emitting ions. The ratio between the intensity of emission from 5D1 to emission from 5D0 increases from Ca
over Sr (Fig. 4) to Ba. This can be related to the lowering of the phonon energy with increasing relative mass of the heavy metal ions
from Ca to Ba in the host lattice.
The excitation spectra of BaF2:Eu nanocrystals co-doped with
different concentrations of Gd3+ions monitored at kem = 592 nm
(5D0 ? 7F1) consist of groups of sharp lines, which can be assigned
to 7F0 ? 5H6 (320 nm), 7F0 ? 5D4 (363 nm), 7F0 ? 5G2 (378 nm),
7
F0 ? 5L6 (397 nm) and 7F0 ? 5D2 (467 nm) ff transition of the
Eu3+ ion (Fig. 5). In materials doped with both, Eu3+ and Gd3+, an
energy transfer between Gd3+ and Eu3+ can occur. Indeed, in the
excitation spectra alongside the characteristic Eu3+ transitions
the 8S7/2 ? 6PJ (310 nm) and 8S7/2 ? 6IJ (272 nm) transitions of
Gd3+ occur. With increasing Gd3+ concentration, the relative intensity between the Gd3+ and Eu3+ transitions changes. The intensity
ratio between the 8S7/2 ? 6IJ (Gd3+) and 7F0 ? 5L6 (Eu3+) transition
rst increase with higher Gd3+ concentration, reaching the largest
intensity ratio for nanoparticles with a concentration of BaF2:Eu
1%, Gd 15%, followed by a decrease with further co-doping degree

339

C. Lorbeer et al. / Optical Materials 34 (2011) 336340

Fig. 2. Representative TEM micrographs of Eu3+ and Gd3+ co-doped alkaline earth uorides.

BaF2:Eu 1 %, Gd 5 %

Intensity / a.u.

Intensity / a.u.

BaF2:Eu 1 %, Gd 5 %
BaF2:Eu 1 %, Gd 10 %
BaF2:Eu 1 %, Gd 15 %

BaF2:Eu 1 %, Gd 10 %

BaF2:Eu 1 %, Gd 15 %

BaF2:Eu 1 %, Gd 20 %

BaF2:Eu 1 %, Gd 20 %
BaF2:Eu 1 %, Gd 25 %

500

BaF2:Eu 1 %, Gd 25 %

550

600

650

700

300

Wavelength / nm

Intensity / a.u.

SrF2: Eu 1 %, Gd 15 %

CaF2: Eu 1 %, Gd 15 %

400

500

400

450

500

550

Wavelength / nm

Fig. 3. Room temperature emission spectra of Eu3+ and Gd3+ co-doped BaF2
7
nanoparticles excited at kex = 393 nm (5L6
F0,1).

300

350

600

700

Wavelength / nm
Fig. 4. Room temperature excitation spectra (left) monitored at kem = 592 nm
7
(5D0 ? 7F1) and emission spectra (right) excited at kex = 393 nm (5L6
F0,1) of SrF2
and CaF2 co-doped with 1% Eu3+ and 15% Gd3+.

of gadolinium. Thus, a concentration of 15% Gd3+ seems to be most


favorable for an efcient Gd ? Eu energy transfer. A similar tendency was observed earlier for Eu3+ doped GdF3 nanocrystallites
[4]. It is important to notice that neither in the excitation spectra
of BaF2:Eu, Gd nor in the excitation spectra of SrF2:Eu, Gd and CaF2:Eu, Gd (Fig. 4) any traces of a EuO charge transfer band could
be detected documenting the potential of our synthesis method.

Fig. 5. Room temperature excitation spectra Eu3+ and Gd3+ co-doped BaF2
nanoparticles monitored at kem = 592 nm (5D0 ? 7F1).

Nanouorides synthesized using conventional synthesis routes like


hydrothermal synthesis very often contain impurities of oxide or
OH, which lead to emission quenching [2527].
Thus, the lifetimes of the europium and gadolinium co-doped
alkaline earth uorides prepared by microwave synthesis from
the ionic liquid [C4mim][BF4] were studied. A decay time of
17 ms was determined for the BaF2 sample co-doped with 5%
Gd3+ excited both directly into europium (kex = 393 nm) and into
gadolinium (kex = 272 nm). For larger gadolinium concentrations
ranging from 10% to 25%, a lifetime of 15.5 ms was obtained. This
decrease could result from a change in symmetry due to the more
considerable need for charge compensation when co-doping more
than 5% of gadolinium, as also apparent in the emission spectrum.
The decay times calculated for SrF2 and CaF2 co-doped with 1% Eu3+
and 15% Gd3+ were somewhat longer with 19.5 ms and 19 ms,
respectively. It is expected that the higher the relative mass of
the metal (alkaline earth cation) in the host structure the lower
the degree of non-radiative relaxation, thus implying a longer lifetime in BaF2. However, in case of alkaline earth uorides this picture is more complex: As a result of the different sizes of the
metal ion of Ca, Sr and Ba with respect to the co-doped lanthanide
ion, different charge compensating processes, thus defect clustering, occur leading to altered site symmetries of the optically active
rare earth ions.
It is well known, that in uoride matrices doped with Gd3+ and
3+
Eu quantum cutting can be observed with quantum efciencies

1,5

BaF2:Gd, Eu

QY=172 %
6

ex. 202 nm ( G J)
6

ex. 305 nm ( PJ )

1,0
0,5
0,0
500

550

600

650

Wavelength / nm

700

2,0

SrF2:Gd, Eu

QY=191 %

1,5
1,0
0,5
0,0
500

550

600

650

Normalized intensity

2,0

Normalized intensity

C. Lorbeer et al. / Optical Materials 34 (2011) 336340

Normalized intensity

340

2,0

CaF 2:Gd, Eu

1,5
1,0
0,5
0,0

700

500

Wavelength / nm

Fig. 6. Emission spectra of 1% Eu3+ and 15% Gd3+ co-doped alkaline earth uorides excited via 6GJ
cutting abilities.

close to 200% [28]. For BaF2:Gd 1%, Eu 1% bulk materials, downconversion with a quantum yield of 194% was reported [29]. In
case of our nanomaterials we do not observe any energy transfer
from Gd3+ to Eu3+ under VUV excitation for the same low concentration of active centers, suggesting a different rare earth ion distribution in the nanoscale material. (For the bulk, the lanthanide ion
clustering tendency was already observed with 1% co-doping [29].)
Our study implies that the distribution of Ln3+ ions at low dopant
concentration (up to 5%, dominating europium transitions in excitation) could be more random in nanomaterials, thus hampering
Ln3+Ln3+ interactions. Nevertheless, for particles with active ion
concentrations starting from 10%, we observed efcient GdEu energy transfer. The efciency of the quantum cutting process was
estimated from the different intensity ratios of 5D1,2 and 5D0 emis8
sion under excitation at 202 nm (6GJ
S transition) and 305 nm
8
(6PJ
S transition) of gadolinium according to the procedure described by Wegh et al. [30]. As the excitation spectra of europium
and gadolinium co-doped BaF2 suggested that the most efcient
EuGd energy transfer occurs for a Gd3+ concentration of 15%
(and 1% Eu3+) we have determined the quantum cutting efciency
for all co-doped samples in the alkaline earth uoride series at that
concentration (Fig. 6, every spectrum is normalized on the emission from the 5D1 level). Indeed, down-conversion is very effective
and the efciencies were calculated to 170% for BaF2, 190% for SrF2
and 150% for CaF2 doped with 1% Eu3+ and 15% Gd3+.
4. Conclusion
Microwave assisted synthesis employing the ionic liquid
[C4mim][BF4] as the reaction medium and reaction partner is a
new and easy synthetic route to pure, oxygen-free Eu3+ and Gd3+
co-doped alkaline earth uorides. Particle sizes below 10 nm for
the co-doped BaF2 and below 20 nm for CaF2 and SrF2 could be
achieved. It was shown for BaF2, that these small nanoparticles
aggregate with increasing lanthanide dopant concentration from
plates to spheres. Variation of the Gd3+ dopant concentration also
has a strong inuence on the optical properties and the Eu3+
Gd3+ energy transfer. A doping degree of 15% Gd3+ yielded the best
results and we were able to obtain highly efcient quantum cutters
with 190% efciency for strontium uoride, 170% for barium uoride and 150% for calcium uoride, respectively.
Acknowledgments
AVM would like to acknowledge support from the European
Research Council through an ERC Starting Grant (EMIL, Contract

QY=153 %

550

600

650

700

Wavelength / nm
8

S (202 nm) and 6PJ

S (305 nm) of gadolinium to check for quantum

no. 200475), the Fonds der Chemischen Industrie through a Dozentenstipendium for A.-V.M. as well as a Doktorandenstipendium for
C.L. and DESY (proposal no. II-20090181). E.Z. and J.C. would like to
thank for support from grant POIG.01.01.02-02-006/09 of the Minister of Science and Higher Education (Poland).
References
[1] G. Liu, B. Jacquier, Spectroscopic Properties of Rare Earths in Optical Materials,
Springer, Heidelberg, Germany, 2008.
[2] M.R. Ranade, A. Navrotsky, H.Z. Zhang, J.F. Baneld, S.H. Elder, A. Zaban, P.H.
Borse, S.K. Kulkarni, G.S. Doran, H.J. Whiteld, Proc. Nat. Acad. Sci. 99 (2002)
6476.
[3] K.A. Gschneider, J. Bnzli, V.K. Pecharsky, Handbook on the physics and
chemistry of rare earths, vol.37, Elsevier, Amsterdam, The Netherlands, 2007.
[4] C. Lorbeer, J. Cybinska, A.-V. Mudring, Chem. Commun. 46 (2010) 571.
[5] C. Lorbeer, J. Cybinska, A.-V. Mudring, Cryst. Growth Des. 11 (2011) 1040.
[6] S.V. Kuznetsov, V.V. Osiko, E.A. Tkatchenko, P.P. Fedorov, Russ. Chem. Rev. 75
(2006) 1065.
[7] A. Bednarkiewicz, A. Mech, M. Karbowiak, J. Lumin. 114 (2005) 247.
[8] F. Wang, X. Fan, D. Pi, M. Wang, Solid State Commun. 133 (2005) 775.
[9] C.M. Bender, J.M. Burlitch, D. Barber, C. Pollock, Chem. Mater. 12 (2000) 1969.
[10] M. Cao, C. Hu, E. Wang, J. Am. Chem. Soc. 125 (2003) 11196.
[11] Y. Jin, W. Qin, J. Zhang, J. Fluorine Chem. 129 (2008) 515.
[12] D. Chen, Y. Yu, F. Huang, P. Huang, A. Yang, Y. Wang, J. Am. Chem. Soc. 132
(2010) 9976.
[13] W.L. Bragg, The analysis of crystals by the X-ray spectrometer, Proc. Roy. Soc.
Lond., Ser. A: Math. Phys. Sci. 89 (1914) 468.
[14] M. Bouffard, J.P. Jouart, M.-F. Joubert, Opt. Mater. 14 (2000) 73.
[15] B.P. Sobolev, A.M. Golubev, L.P. Otroshchenk, V.N. Molchanov, R.M. Zakalyukin,
E.A. Ryzhova, P. Herrero, Cryst. Rep. 48 (2003) 944.
[16] J.-P.R. Wells, T. Dean, R.J. Reeves, J. Lumin. 96 (2002) 239.
[17] J.P. Jouart, C. Bissieux, G. Mary, J. Lumin. 37 (1987) 159.
[18] A. Gektin, N. Shiran, V. Nesterkina, Y. Boyarintseva, V. Baumer, G. Stryganyuk,
K. Shimamura, E. Villora, J. Lumin. 129 (2009) 1538.
[19] P. Wasserscheid, T. Welton, Ionic liquids in Synthesis, 2nd ed., Wiley-VCH,
Weinheim, Germany, 2008.
[20] J. Rodriguez-Carvajal, Physica B 192 (1993) 55.
[21] T. Roisnel, J. Rodriguez-Carvajal, WinPLOTR: a Windows tool for powder
diffraction patterns analysis, in: R. Delhez, E.J. Mittenmeijer (Eds.), Materials
Science Forum, Proceedings of the Seventh European Powder Diffraction
Conference (EPDIC 7), 118, 2000.
[22] A.M. Golubev, A.K. Ivanov-Shits, V.I. Simonov, B.P. Sobolev, N.I. Sorokina, P.P.
Fedorov, Solid State Ionics 37 (1990) 115.
[23] A.M. Golubev, B.P. Sobolev, V.I. Simonov, Kristallograya 30 (1985) 314.
[24] G. Zimmerer, Rad. Measur. 42 (2007) 859;
G. Zimmerer, J. Lumin. 106 (2006) 1 and http://hasylab.desy.de/facilities/
doris_iii/beamlines/i_superlumi/experimental_station/index_eng.html
(accessed 02.15.11).
[25] I.A. de Crcer, P. Herrero, A.R. Landa-Cnovas, B. Sobolev, Appl. Phys. Lett. 87
(2005) 053105.
[26] P. Ptacek, H. Schfer, K. Kmpe, M. Haase, Adv. Funct. Mater. 17 (2007) 3843.
[27] L. Zhu, X. Liu, J. Meng, X. Cao, Cryst. Growth Des. 7 (2007) 2505.
[28] R.T. Wegh, H. Donker, U.D. Oskam, A. Meijerink, Science 283 (1999) 663.
[29] B. Liu, Y. Chen, C. Shi, H. Tang, Y. Tao, J. Lumin. 101 (2003) 155.
[30] R.T. Wegh, H. Donker, U.D. Oskam, A. Meijerink, J. Lumin. 82 (1999) 93.

S-ar putea să vă placă și