Download as pdf or txt
Download as pdf or txt
You are on page 1of 34

Journal of Nanoparticle Research

Influence of the synthesis conditions of Y0.9Dy0.1VO4 and silica-coated


Y0.9Dy0.1VO4 nanophosphors on the powder morphology and luminescence emission
intensity
--Manuscript Draft--

Manuscript Number: NANO-D-18-01915R1

Full Title: Influence of the synthesis conditions of Y0.9Dy0.1VO4 and silica-coated


Y0.9Dy0.1VO4 nanophosphors on the powder morphology and luminescence emission
intensity

Article Type: Original research

Keywords: nanocomposites; Silica coating; optical properties; Photoluminescence

Abstract: Y0.9Dy0.1VO4 nanophosphors were prepared by a sol-gel process and a


hydrothermal synthesis at different pH values of the reaction medium. Samples
prepared by hydrothermal synthesis at pH = 2 were also coated with silica. All obtained
samples were characterized by X-ray diffraction (XRD), Fourier-transform infrared
spectroscopy (FTIR), transmission electron microscopy (TEM), micro-Raman and
micro-photoluminescence (PL). XRD patterns of Y0.9Dy0.1VO4 samples show
diffraction maxima which can be indexed to a tetragonal symmetry of space group
I41/amd with Z = 4, compatible with a zircon-type structure. A higher intensity of the
XRD maxima is observed in patterns of samples prepared by the sol-gel method at
acid pH. TEM images reveal that the particle morphology and size, ranging from 20 to
45 nm in the different nanophosphors, depend on the preparation method and the
synthesis conditions. Raman and FTIR spectra of the samples have been correlated
with XRD results. An increased thickness of the silica coating is observed in TEM
images as the reaction time with tetraethyl orthosilicate (TEOS) increases. The
presence of silica is also evidenced by FTIR absorption spectra. The different
synthesis routes modify the overall PL intensity, as well as the relative weight of the
dominant blue and yellow emissions characteristics of Dy3+ ions. An enhanced PL
emission intensity has been found in samples coated with silica, which can be
attributed to a reduction of surface defects.

Response to Reviewers: Influence of the synthesis conditions of Y0.9Dy0.1VO4 and silica-coated


Y0.9Dy0.1VO4 nanophosphors on the powder morphology and luminescence emission
intensity, by L. Alcaraz, J. Isasi, and C. Díaz - Guerra, (NANO-D-18-01915)
We would like to thank the referees for their positive comments.

Response to reviewer #1
1. The temperature should be given in Celsius unit as well in whole manuscript.
As suggested by the referee, temperatures have been written in degrees Celsius
throughout the revised manuscript.

2. How did the authors calculate the particle size in using XRD? If they use "Checkcell
software", more explanation is required.
It is a mistake we made in the original manuscript. Checkcell software was used to
calculate the cell parameters (a and c) as well as the unit cell volumes (V), not the
particle size. Caption to Table 1 has been modified in the revised version of our work.

3. Table 1 is not clear. The authors should clarify what is the (a, c and V).
As requested by the referee, caption to Table 1 has been modified in order to make it
more informative.
“Table 1. Cell parameters (a and c) and unit cell volume (V) calculated for the samples
investigated.”

4. For size measurement in TEM images, it seems just few particles have been
measured, so the reported size is not accurate.
To estimate the average particle size, a minimum of 80 particles were measured in
different TEM micrographs for each sample. Due to the slight dispersion found in the
average particle size, the authors consider that the results are representative for the
samples investigated.

Powered by Editorial Manager® and ProduXion Manager® from Aries Systems Corporation
Response to reviewer #2

1. Why pH=2 nanophosphor (by hydrothermal method) is chosen for silica coating
Spherical nanoparticles of smaller size exhibit higher surface to volume ratios. Among
the samples investigated in this work, this is the morphology characteristic of sample
HT-2, which explains why it was chosen to investigate the effects of silica coating.
A sentence has been added in the revised manuscript (page 10) to explain this.
“Since nanoparticles of the latter sample show a smaller size and a rather uniform
shape, silica-coating effects were investigated for this material”.

2. Clarification needed for Dy3+ intraionic transition.


3. Peak positions are not influenced by synthetic route - Give reason.
As explained in the original version of our work (page 14), “all emission lines observed
are in good agreement with the optical transitions reported for Dy3+ in a D2d symmetry
site, as expected for Dy3+ substituting Y3+ in the YVO4 lattice [53, 54], and can be
attributed to Dy3+ 4f-4f intraionic transitions”. Neither the synthesis route followed in
order to obtain our nanophosphors, nor the acidity (pH value) of the reaction medium,
influence the peak position of the observed PL emissions. Peak positions mainly
depend on the crystal-field splitting of the energy levels of Dy3+ in the orthovanadate
host, which in turn depends on the site symmetry of the rare earth. According to our
XRD data, the crystal structure of all obtained samples is the same. Little variations in
the peak positions are thus found when PL spectra of the different powders are
compared, which also evidences that Dy ions occupy similar sites in all the
nanophosphors investigated.
A new paragraph has been added (page 16) to the revised manuscript in order to
make these points clear.
“…which strongly suggests that Dy ions occupy similar sites in all the nanophosphors
investigated, since peak positions mainly depend on the crystal-field splitting of the
energy levels of Dy3+ in the orthovanadate host, which in turn depends on the site
symmetry of the rare earth. This is also supported by our XRD results, showing the
same crystal structure for all the investigated samples.”

4. Reason for enhanced PL emission for silica coated materials.


As stated in the original manuscript (page 18) the mentioned enhancement “can be
attributed to a reduction of surface defects characteristic of the smaller nanoparticles.
Such defects are considered to provide non-radiative recombination channels for
electrons and holes, leading to a reduced quantum yield of the nanophosphors.”
Defects at surfaces provide new states for electrons and holes, which may alter their
motion, lifetime, and transition energies. Precisely, dangling bonds at a semiconductor
or insulating surfaces frequently give rise to electronic states within the band-gap.
These mid-gap states fill up to the Fermi level with electrons that originate in the bulk of
the material. The accumulation of charge at the surface creates an electric field – a
depletion region – that leads to bending of the valence and conduction band edges.
According to a simple dead layer model, electron–hole pairs generated in this region
are swept apart by the electric field, facilitating rapid non-radiative recombination.
Moreover, dangling bonds often act as getters for impurities as well. Although the PL
response is influenced by several mechanisms, it is generally found that large PL
signals correlate with good quality surfaces. In particular, surface recombination
velocity and surface band bending play a key role in this context, as just mentioned.
Because surface recombination is usually non-radiative and band bending can lead to
the formation of the mentioned depletion region where PL is effectively quenched, both
of these phenomena tend to reduce the PL intensity. In the present case, silica coating
effectively passivates surface dangling bonds and prevents the formation of the
described depletion layer.
A new paragraph has been added (page 18) to the revised manuscript in order to
include this explanation.
“In fact, defects at surfaces provide new states for electrons and holes, which may alter
their motion, lifetime, and transition energies. Precisely, dangling bonds at a
semiconductor or insulating surfaces frequently give rise to electronic states within the
band-gap which fill up to the Fermi level with electrons that originate in the bulk. The
accumulation of charge at the surface creates an electric field – a depletion region –
that leads to bending of the valence and conduction band edges. According to a simple
dead layer model, electron–hole pairs generated in this region are swept apart by the

Powered by Editorial Manager® and ProduXion Manager® from Aries Systems Corporation
electric field, facilitating rapid non-radiative recombination. Moreover, dangling bonds
often act as getters for impurities as well. Although the PL response is influenced by
several mechanisms, it is generally found that large PL signals correlate with good
quality surfaces. In particular, surface recombination velocity and surface band
bending play a key role in this context, as just mentioned. Since surface recombination
is usually non-radiative and band bending can lead to the formation of the depletion
region where PL is effectively quenched, both of these phenomena tend to reduce the
PL intensity. In the present case, it appears that silica coating effectively passivates
surface dangling bonds and prevents the formation of the described depletion layer.”

Response to reviewer #3
1. I have read the manuscript in detail but fail to understand why author had mentioned
under introduction section that they are not aware of any previous study done in area
related to their work. This statement puts question mark on the originality of the work. It
would have been better, if the author would have mentioned that either no work has
been reported or if the work is already reported in literature, then how that is different
from their own research work. Plz clarify this statement.

The referee states the authors mentioned in the Introduction section of the original
manuscript “that they are not aware of any previous study done in area related to their
work.” This is not correct. Actually, the authors clearly explain that Er and Eu-doped
vanadates have been frequently investigated, which is not the case for other rare earth
dopants, such as Dy. Moreover, they also clearly explain the aim of their work, to carry
out a comparative assessment of the influence of synthesis methods and conditions on
the particle morphology, structure and luminescence properties of Y0.9Dy0.1VO4 and
silica-coated Y0.9Dy0.1VO4 nanophosphors. This information was already contained
in page 4 of the original manuscript:

“The synthesis and study of Ln3+:YVO4 nanophosphors are frequently described in


the literature. However, yttrium orthovanadate activated with different concentrations of
trivalent dysprosium has been studied to a much lesser extent than Eu or Er-doped
YVO4. In particular, to the best of our knowledge, we are not aware of previous works
reporting a comparative assessment of the influence of synthesis methods and
conditions on the particle morphology, structure and luminescence properties of
Y0.9Dy0.1VO4 and silica-coated Y0.9Dy0.1VO4 nanophosphors, which is the aim of
the present work.”

2. The language needs major improvement in whole manuscript. For e.g. Text written
under experimental section (especially under Hydrothermal Synthesis) needs major
improvement in language.

The whole manuscript has been thoroughly revised by an English native speaker.

Additional Information:

Question Response

Scientific Justification (Available to Attached please find the revised version of our manuscript “Influence of the synthesis
Reviewers) conditions of Y0.9Dy0.1VO4 and silica-coated Y0.9Dy0.1VO4 nanophosphors on the
powder morphology and luminescence emission intensity”, by L. Alcaraz, J. Isasi and
C. Díaz-Guerra (NANO-D-18-01915), to be considered for publication in Journal of
Nanoparticle Research.
We would like to thank the reviewers for their comments and suggestions, which
helped us to increase the quality and clarity of our work. Such comments have been
addressed on a point by point basis.
As requested by the editorial office, we have included the size range of the
nanoparticles in the abstract of this revised work. The manuscript has also been
modified to list the references in alphabetical order and cite them in the text using the
(author, year) format.
We sincerely hope you will consider this revised version of our work suitable for
publication in Journal of Nanoparticle Research.

Powered by Editorial Manager® and ProduXion Manager® from Aries Systems Corporation
Manuscript_revised version Click here to access/download;Manuscript;Alcaraz et
al_NANO-D-18-01915_revised version.doc
Click here to view linked References

Influence of the synthesis conditions of Y0.9Dy0.1VO4 and silica-coated Y0.9Dy0.1VO4


nanophosphors on the powder morphology and luminescence emission intensity

L. Alcaraz1, J. Isasi1, *, C. Díaz-Guerra2

1
Departamento de Química Inorgánica I, Facultad de Ciencias Químicas, Universidad
Complutense de Madrid, Ciudad Universitaria s/n, 28040 Madrid, Spain.

2
Departamento de Física de Materiales, Facultad de Ciencias Físicas, Universidad
Complutense de Madrid, Ciudad Universitaria s/n, 28040 Madrid, Spain.

authors e-mails: lorena.alcaraz@ucm.es; isasi@ucm.es; cdiazgue@ucm.es

Keywords: Nanocomposites; Silica coating; Optical properties; Photoluminescence

* Corresponding author
e-mail: isasi@ucm.es
Phone: +34 913945215
1
Abstract

Y0.9Dy0.1VO4 nanophosphors were prepared by a sol-gel process and a hydrothermal synthesis


at different pH values of the reaction medium. Samples prepared by hydrothermal synthesis at
pH = 2 were also coated with silica. All obtained samples were characterized by X-ray
diffraction (XRD), Fourier-transform infrared spectroscopy (FTIR), transmission electron
microscopy (TEM), micro-Raman and micro-photoluminescence (PL). XRD patterns of
Y0.9Dy0.1VO4 samples show diffraction maxima which can be indexed to a tetragonal
symmetry of space group I41/amd with Z = 4, compatible with a zircon-type structure. A
higher intensity of the XRD maxima is observed in patterns of samples prepared by the sol-
gel method at acid pH. TEM images reveal that the particle morphology and size, ranging
from 20 to 45 nm in the different nanophosphors, depend on the preparation method and the
synthesis conditions. Raman and FTIR spectra of the samples have been correlated with XRD
results. An increased thickness of the silica coating is observed in TEM images as the reaction
time with tetraethyl orthosilicate (TEOS) increases. The presence of silica is also evidenced
by FTIR absorption spectra. The different synthesis routes modify the overall PL intensity, as
well as the relative weight of the dominant blue and yellow emissions characteristics of Dy3+
ions. An enhanced PL emission intensity has been found in samples coated with silica, which
can be attributed to a reduction of surface defects.

2
Introduction

As an important family of luminescent materials, lanthanide doped orthovanadates


(Ln3+:YVO4) have been paid attention to because of their excellent properties, e.g., high
thermal and chemical stability, low cost, high quantum efficiency, high absorption of VO43- in
the UV range, moderate phonon energy and an exceptional optical damage threshold (Assefa
et al. 2004; Moon et al. 2007; Jang et al. 2009). In fact, as a host lattice for rare earth ions,
YVO4 can be excited under UV light and the photon energy efficiently transferred from the
excited vanadate groups to the trivalent rare earth ions (Xu et al. 2007). In particular, Dy3+
ions have been the subject of current research as single dopants as well as activators in
different aluminates (B. Liu, C.S. Shi 2005; Omkaram and Buddhudu 2009; Vijayakumar and
Marimuthu 2015) and silicates hosts (Y.L. Liu, B.F.Lei 2005; Kuang et al. 2006; Liu et al.
2007; Chen et al. 2009; You et al. 2011) due to their high quantum efficiency, providing
appropriate color rendering index and correlated color temperature (Shrivastava et al. 2016).
The role of Dy3+ ions as sensitizers in a number ofcrystalline phosphors has also been
addressed (Li et al. 2008; Xie et al. 2015). It is known that luminescence of Dy3+ ions in the
visible spectral range is mainly a consequence of two different groups of radiative transitions
from the excited 4F9/2 level to the 6H15/2 and 6H13/2 levels, which respectively give rise to
emission lines in the blue (470–500 nm) and yellow (560–590 nm) ranges (Gu et al. 2004; Su
and Yan 2005; Wang et al. 2009a). 4F9/2→6H13/2 transitions are dominant only when Dy3+ ions
are located at low-symmetry sites with no inversion center. Nevertheless, it should be noted
that the optical properties of Ln3+:YVO4 materials - in particular the intensity and spectral
distribution of the PL emission (Ray et al. 2009; Shinde et al. 2012; Yuan et al. 2013; Alcaraz
et al. 2014, 2015a; b) - are influenced not only by the specific host lattice and the crystal
environment of the activator ions, but also by the particle size, morphology and degree of
agglomeration (Alcaraz et al. 2014, 2015a; b, 2016b).

A variety of preparation methods based on aqueous and non-aqueous routes have been
reported in the literature for the synthesis of Ln3+:YVO4 nanophosphors, such as solid state
reactions (Yan and Su 2005; Wu et al. 2007), chemical co-precipitation (Kuang et al. 2006),
sol-gel processes (Bao et al. 2008) or hydrothermal synthesis (Song et al. 2012).
Unfortunately, in the first two cases the treatments at high temperature lead to an increase of
the particle size, which also results in the obtaining of agglomerated powders. On the
contrary, the development of sol-gel or hydrothermal processes currently allows the synthesis

3
of non-agglomerated nanoparticles with more uniform morphologies using relatively low
temperatures. Such processes have already been tested by the authors of this work for the
synthesis of compounds doped with different rare earth ions (Yan and Su 2005; Wu et al.
2007; Bao et al. 2008; Sun et al. 2010) and several stoichiometries (Alcaraz et al. 2015c,
2016a; Isasi et al. 2018).

In some circumstances, during the different steps of the synthesis process followed in order to
obtain samples with high luminescence emission intensity, the formation of surface defects
may take place. Such defects may lead to a decreased quantum efficiency of the synthesized
nanophosphors due to enhanced non-radiative recombination. In order to minimize this
problem, core-shell structures have been sometimes used (Ocaña et al. 2011; Liu et al. 2015).
In particular, silica has been extensively used as an efficient coating because of its high
chemical stability against oxidation, optical transparency and easily controllable shell
thickness (Ocaña et al. 2011; Liu et al. 2015). The synthesis and study of Ln3+:YVO4
nanophosphors are frequently described in the literature. However, yttrium orthovanadate
activated with different concentrations of trivalent dysprosium has been studied to a much
lesser extent than Eu or Er-doped YVO4. In particular, to the best of our knowledge, we are
not aware of previous works reporting a comparative assessment of the influence of synthesis
methods and conditions on the particle morphology, structure and luminescence properties of
Y0.9Dy0.1VO4 and silica-coated Y0.9Dy0.1VO4 nanophosphors, which is the aim of the present
work.

Experimental

Synthesis

Y0.9Dy0.1VO4 samples were prepared by sol-gel and hydrothermal synthesis at pH = 2 and 9.


In addition, some silica-coated Y0.9Dy0.1VO4 samples were synthesized by using the Stöber
method (Stober and Fink 1968). Yttrium nitrate hexahydrate (Y(NO3)3·6H2O, 99.9% Strem
Chemicals), dysprosium nitrate (Dy(NO3)3·H2O, 99.9% Strem Chemicals) and ammonium
vanadate (NH4VO3, 99% Aldrich, analytical grade) were used as starting materials, while
citric acid (CA, 99.5% Panreac) and ethylene glycol, EG, (99% Sigma Aldrich) were used as
chelating and polymeric agents, respectively. Finally, tetraethyl orthosilicate (TEOS, Merck)
was used for the coating of Y0.9Dy0.1VO4 samples.

4
Sol-gel synthesis

0.5 g of yttrium nitrate hexahydrate, 0.05 g of dysprosium nitrate and 0.17 g of ammonium
vanadate were dissolved in 50 mL of deionized water. The pH value was adjusted to 2 by the
addition of a CA solution, according to a 1:1 molar ratio with respect to vanadium. In order to
investigate the effects of the pH variation on the structure, powder morphology and
luminescence emission intensity, Y0.9Dy0.1VO4 samples were also prepared at pH 9 by the
addition of NH3 (25% Sigma Aldrich). Independently of the reaction medium used, 2 mL of
EG were slowly added to the respective solutions to generate weak interactions between the
particles during crystallization, leading to the obtaining of dispersed nanoparticles with a
better luminescence emission intensity (Hassanzadeh-Tabrizi 2012). When dissolution was
complete, the temperature was increased to 140 ºC to remove water excess until gels were
formed. These gels were aged for 4 days and dried at 250 ºC for 1 hour. The obtained
precursor powders were treated at 750 ºC for 4 hours obtaining samples of white color which
are named SG-2 and SG-9.

Hydrothermal synthesis

0.25 g of yttrium nitrate hexahydrate, 0.025 g of dysprosium nitrate and 0.085 g of


ammonium vanadate were dissolved in 30 mL of deionized water. The pH values of the
solutions (2 and 9) were adjusted in the same way described for the sol-gel samples. The
resulting mixtures were then stirred during 20 min and subsequently placed in a Teflon-lined
stainless-steel autoclave. In all cases, the autoclave was heated at 130 ºC for 24 hours. After
cooling down to room temperature, the obtained precipitates were dried at 100 ºC for 1 hour.
These samples were named HT-2 and HT-9.

In order to obtain silica-coated samples, 0.025 g of HT-2 material were dispersed in a (1:3)
deionized water:ethanol mixture. Then, 5 mL of ammonia (32%) were added and the final
solution was sonicated. After complete homogenization of the mixture, 1 mL of TEOS was
slowly added under constant magnetic stirring. The reaction time with TEOS was set to either
6 or 18 h, in order to assess the effect of this parameter on the silica coating thickness. The
obtained powders were separated by centrifugation and washed twice with a solution of
deionized water and ethanol. The obtained samples were respectively called HT-2_TEOS_6
and HT-2_TEOS_18.

5
Material characterization

XRD patterns of the synthesized samples were recorded on a Philips X'Pert PRO MPD
diffractometer using Cu Kα1 radiation with a step size of 0.02º (2) and a step time of 7.9 s.
FTIR spectra were recorded on a Prestige-21 Fourier Transform spectrophotometer using the
KBr pellet technique within the (2000–400) cm-1 range. The morphological characterization
of the samples was carried out by transmission electron microscopy (TEM) using a JEOL
JEM 2100 instrument operating at 200 kV. For TEM observations, the powders were
dispersed in n-butanol and drops of the corresponding suspensions were deposited on carbon-
coated copper grids. TEM-energy dispersive spectroscopy (EDS) analyses were carried out by
using an OXFORD INCA instrument. Micro-Raman and photoluminescence (PL)
measurements were carried out at room temperature in a Horiba Jovin-Ybon LabRAM HR800
system. The samples were excited by a 325 nm He–Cd laser on a confocal Olympus BX 41
microscope with a 40x objective. The spectral resolution of the system used was 1.5 cm -1 for
micro-Raman measurements and 0.1 nm for PL measurements. The CIE (Commission
Internationale de l’Eclairage) color coordinates were calculated using Color Calculator 6.03
software provided by Osram Sylvania.

Results and discussion

X-ray diffraction (XRD)

Fig. 1 shows XRD patterns of the Y0.9Dy0.1VO4 samples prepared by both sol-gel and
hydrothermal synthesis at pH = 2 and 9. All observed reflexions can be indexed to a
tetragonal symmetry (I41/amd space group and Z = 4) [JCPDS file 17-0341] compatible with
a zircon-type structure.

6
Fig. 1 XRD patterns of the Y0.9Dy0.1VO4 samples synthesized by the hydrothermal (HT) and
sol-gel (SG) processes at pH 2 and 9

A higher intensity of the XRD maxima is observed for samples prepared by the sol-gel
method, which can be attributed to the high temperature (750 ºC) at which these samples were
treated during the synthesis process. It can also be appreciated how the acidic medium gives
rise to more intense reflections and therefore more crystalline samples. The X-ray diffraction
patterns of HT-2 and HT-9 samples indicate a lower crystallinity with a slight variation in the
intensity of the diffraction maxima. The lattice parameters and cell volumes of Y0.9Dy0.1VO4
samples calculated with Checkcell software are shown in Table 1. Lattice parameters of
samples prepared by sol-gel are somewhat larger than those obtained by hydrothermal

7
synthesis. These results are in good agreement with our studies of similar systems doped with
other lanthanide ions (Alcaraz et al. 2015b, 2016b).

Table 1. Cell parameters (a and c) and unit cell volume (V) calculated for the samples
investigated.

Samples a = b (Å) c (Å) V (Å3)

SG-2 7.121(3) 6.266(5) 317.74

SG-9 7.120(5) 6.265(4) 317.61

HT-2 7.115(8) 6.261(4) 317.15

HT-9 7.118(6) 6.263(7) 317.32

Fourier transform infrared spectroscopy (FTIR).

Fig. 2 shows FTIR spectra of Y0.9Dy0.1VO4 and silica-coated samples. Absorption bands are

observed in the (2000–400) cm-1 range.

The most intense band (centred at about 820 cm-1) can be attributed to the antisymmetric

stretching mode of VO43- units (aimay Lin-Vien, Norman B. Colthup 1991; Wang et al. 2006;

Liang et al. 2012). The bands present at around 450 cm-1 can be assigned to the vibration of

Y/Dy-O bonds (Yu et al. 2002; Ningthoujam et al. 2009; Grandhe et al. 2012). It is noted that

the absorption band at 450 cm-1 observed in the FTIR spectra of silica-coated samples is more

intense, probably due to the overlapping with the Si-O bond bending vibration that appears

centred between 468 and 453 cm-1 (Wang et al. 2009b; Shui et al. 2013; Lin et al. 2014).

Bands between 1100 and 1220 cm-1 can be attributed to Si-O-Si and O-Si-O stretching

vibration modes (Wang et al. 2009b; Shui et al. 2013). The absorption band at around 1640

8
cm-1 can be attributed to the H2O bending vibrations of water adsorbed in the KBr used in the

preparation of the pellets.

Fig. 2 FTIR spectra of Y0.9Dy0.1VO4 and silica-coated samples

Transmission electron microscopy (TEM)

Fig. 3 shows TEM images of Y0.9Dy0.1VO4 samples prepared by sol-gel and hydrothermal
synthesis at acid and basic pH, as well as images of the silica-coated nanophosphors. The
cation compositions determined by EDS microanalysis in different areas of the samples (see
9
Table 2) are in good agreement with the stoichiometric composition. TEM images of
Y0.9Dy0.1VO4 samples prepared by sol-gel show spherical particles with an average diameter
of 30 nm (SG-2) and 42 nm (SG-9). For Y0.9Dy0.1VO4 samples prepared by hydrothermal
synthesis, TEM images show particles of elongated shape, with a mean diameter of 20 nm
and lengths of about 45 nm in the case of the HT-9 material, while TEM micrographs of HT-2
sample show spherical particles with an average size of 25 nm. Since nanoparticles of the
latter sample show a smaller size and a rather uniform shape, silica-coating effects were
investigated for this material. Agglomerated nanoparticles coated with a silica shell can be
observed in TEM images of HT-2_TEOS_6 and HT-2_TEOS_18 samples. The thickness of
this shell is approximately 100 nm for sample HT-2_TEOS_6 and 250 nm for the HT-
2_TEOS_18 sample, revealing that and increased TEOS reaction time results in an increased
silica coating thickness. These results are in good agreement with those previously reported
for materials with similar stoichiometry (Ray et al. 2009; Alcaraz et al. 2014; Alcaraz and
Isasi 2017) as well as Li1+xV3O8 samples (Shui et al. 2013) and Ag-SnO2 composite powders
(Lin et al. 2014).

10
Fig. 3 TEM images of the Y0.9Dy0.1VO4 samples prepared by sol-gel: SG-2 (a), SG-9 (b) and
by hydrothermal synthesis: HT-2(c), HT-9 (d), HT-2_TEOS_6 (e) and HT-2_TEOS_18 (f)

Table 2. Cation compositions (at. %) determined by TEM-EDS microanalysis.

Stoichiometric Sol-gel method Hydrothermal synthesis


Cation at. %
pH = 2 pH = 9 pH = 2 pH = 9
V 50 51.4 50.8 48.1 49.2
Y 45 43.4 44.5 46.2 45.7
Dy 5 5.1 4.8 5.8 5.2

Raman spectroscopy
11
Fig. 4 shows normalized Raman spectra of Y0.9Dy0.1VO4 samples synthesized by the sol-gel
method, while Raman spectra of uncovered and silica-covered samples obtained by the
hydrothermal synthesis are shown in Fig. 5. The positions of the observed bands and their
corresponding assignments are shown in Table 3.

12
Fig. 4 Normalized Raman spectra of .SG-2 and SG-9 samples

Fig. 5 Normalized Raman spectra of HT-2, HT-9, HT-2_TEOS_6 and HT-2_TEOS_18


samples

13
Table 3. Raman frequencies and symmetry assignments for the synthesized samples.

Sample Raman Peaks (cm-1)

Assignment ν2 (B1g) ν 2 (A1g) ν 4 (B2g) ν 3 (B2g) ν 3 (Eg) ν 1 (A1g)

SG-2 261 380 489 816 838 891

SG-9 262 379 --- 817 839 891

HT-2 261 379 488 817 838 892

HT-9 261 379 489 816 838 892

HT-2_TEOS_6 263 380 489 816 838 892

HT-2_TEOS_18 263 380 489 816 838 892

It can be appreciated that the peak positions of the observed bands and their relative
intensities are almost independent of the sample considered. These results are similar to those
obtained in Raman measurements of undoped or rare-earth doped orthovanadates (Miller et
al. 1968; Long et al. 2006; Sadhu et al. 2007; Voron’ko et al. 2009; Manjón et al. 2010;
Alcaraz et al. 2014, 2015b). Briefly, the observed Raman peaks are due to internal modes
corresponding to the vibrations of the oxygen atoms in the VO43− units (ν1–ν4). Raman bands
corresponding to symmetric stretching vibrations (ν1) have been usually reported peaked at
about 891 cm-1, while antisymmetric stretching vibrations (ν3) appear centred at about 838
(Eg) and 817 (B2g) cm-1, respectively. On the other hand, symmetric bending (ν2) vibrations
appear usually peaked near 260 (B1g) and 379 (A1g) cm-1 while antisymmetric bending
vibrations (ν4) give rise to a Raman peak at about 489 cm-1 (Miller et al. 1968; Long et al.
2006; Sadhu et al. 2007; Voron’ko et al. 2009; Manjón et al. 2010). In the present case, a
clear broadening of all the peaks present in Raman spectra of the HT-9 sample can be
appreciated in Fig. 5, evidencing a certain distortion or non-equivalency of vibrations of
VO43- units. These results are in rather good agreement with the crystallinity observed in the
corresponding XRD patterns (see Fig. 1), where it was found that samples prepared by
hydrothermal synthesis show diffraction maxima of lower intensity. It is noted that Raman
spectra of silica-covered samples (see Fig. 5) are coincident with those obtained in the
uncoated samples.

14
Photoluminescence study

A comparison of PL spectra of the uncovered and silica-covered Y0.9Dy0.1VO4 samples is


shown in Fig. 6. All emission lines observed are in good agreement with the optical
transitions reported for Dy3+ in a D2d symmetry site, as expected for Dy3+ substituting Y3+ in
the YVO4 lattice (Cavalli et al. 2002; Faoro et al. 2009), and can be attributed to Dy3+ 4f-4f
intraionic transitions. Peak positions of the Dy3+ emission bands observed in PL spectra of the
investigated samples and their assignments are shown in Table 4.

Fig. 6 Normalized PL spectra of Y0.9Dy0.1VO4 and silica-covered samples

15
Table 4. Peak positions of the Dy3+ emission bands observed in PL spectra of all investigated
samples and their assignments.

Dy-doped. Peak positions ( ± 0.1 nm ) Assignment Comments

450.1, 454.5, 456, 458, 461 4


I15/2 – 6H15/2

2nd group of emissions in


474.1, 477.3, 478.3, 479.9, 481.5, 484.1 4
F9/2 – 6H15/2
terms of PL intensity

537, 539, 540.5, 542, 542.7 4


I15/2 – 6H13/2

568, 569.1, 572.3, 574, 574.5, 575.1, Dominant group of


4
F9/2 – 6H13/2
577.5, 578.2, 578.2, 579, 583 emissions

599.5, 602, 602.3, 605 4


I15/2 – 6H11/2

650, 652, 656.6, 658.8, 660.6, 661.3 4


F9/2 – 6H11/2

666.7, 664, 664.7, 665.3, 666.7, 667.2, 4


G11/2 – 6H9/2
668.7
4
733, 741, 743, 746.5, 747.3, 750, 752.6, F9/2 - 6H9/2 +
6
753.9, 757.8, 760, 763 F11/2
6
771.5 F3/2 - 6H7/2
4
813, 821, 827, 829.7, 836.5, 838.8, F9/2 - 6H7/2 +
6
841.6, 845, 847, 853, 855 F9/2

The position of the emission lines differs in less than 0.1 nm for samples prepared by the two
methods under different pH values, which strongly suggests that Dy ions occupy similar sites
in all the nanophosphors investigated, since peak positions mainly depend on the crystal-field
16
splitting of the energy levels of Dy3+ in the orthovanadate host, which in turn depends on the
site symmetry of the rare earth. This is also supported by our XRD results, showing the same
crystal structure for all the investigated samples. However, some differences were found in
the intensity of the emission observed in PL spectra of the different samples. Although a
quantitative measurement of the quantum efficiency of each material was not carried out,
spectra were measured under the same experimental conditions, which allowed a direct and
reproducible comparison of the PL emission intensity of the different nanophosphors. The
obtained results are shown in Table 5. It was found that the preparation method and synthesis
conditions mainly modify the overall integrated intensity and also influence the relative
weight of several transitions observed in the PL spectra corresponding to different samples.
The highest PL intensity was found in samples prepared by sol-gel and especially in the SG-2
sample. In the case of samples prepared by hydrothermal synthesis, the opposite trend is
observed. These results might be related to the degree of crystallinity and the particle size of
the synthesized nanopowders. Actually, a high degree of crystallinity was observed in the
XRD patterns of samples prepared by sol-gel (see Fig. 1), leading to a high PL emission
intensity, which is in good agreement with previous reports on rare earth-doped vanadates
(Alcaraz et al. 2015b), fluorides (Li et al. 2013) and other complex oxides (Xue and Sun
2014).

Table 5. Relative PL emission intensity of all samples investigated.

Sample Relative PL intensity


SG2 100
SG9 55.5
HT2 6.5
HT9 38.5
HT-2_TEOS_6 10.4
HT-2_TEOS_18 7.5

Regarding silica-coated samples, it has been reported that silica coating of rare earth-doped
inorganic nanoparticles may either improve or decrease the quantum efficiency, lifetime or
17
photostability of the core material due to the interplay of several competing factors (Wang et
al. 2009b; Kai et al. 2010; Luwang et al. 2011). Valuable results have generally been obtained
owing to the reduction of PL surface quenching effects related to structural defects or an
increased surface reactivity (Wang et al. 2009b; Kai et al. 2010). However, in other cases
(Luwang et al. 2011; Liu et al. 2012) detrimental effects were observed, mainly due to the
presence of silanol (SiO-H) groups which act as luminescence quenchers. Besides, the power
of the excitation light may become weak due to scattering and reflection of the silica shell if
this coating is too thick, which may also modify the position of spectral lines and branching
ratios due to reabsorption processes (Luwang et al. 2011; Liu et al. 2012). In the present work,
silica coating was not found to modify the PL spectral distribution of the core nanoparticles
but to induce a luminescence enhancement of the PL emission intensity. Such enhancement is
more noticeable in the case HT-2_TEOS_6 sample, where TEM images (see Fig. 3) reveal the
formation of a homogeneous, 100 nm thick, silica coating. By increasing the reaction time to
18 hours, the silica shell becomes thicker (250 nm, sample HT-2_TEOS_18) and the
luminescence intensity slightly increases. These results can be attributed to a reduction of
surface defects characteristic of the smaller nanoparticles. Such defects are considered to
provide non-radiative recombination channels for electrons and holes, leading to a reduced
quantum yield of the nanophosphors (Luwang et al. 2011). In fact, defects at surfaces provide
new states for electrons and holes, which may alter their motion, lifetime, and transition
energies. Precisely, dangling bonds at a semiconductor or insulating surfaces frequently give
rise to electronic states within the band-gap which fill up to the Fermi level with electrons that
originate in the bulk. The accumulation of charge at the surface creates an electric field – a
depletion region – that leads to bending of the valence and conduction band edges. According
to a simple dead layer model, electron–hole pairs generated in this region are swept apart by
the electric field, facilitating rapid non-radiative recombination. Moreover, dangling bonds
often act as getters for impurities as well. Although the PL response is influenced by several
mechanisms, it is generally found that large PL signals correlate with good quality surfaces.
In particular, surface recombination velocity and surface band bending play a key role in this
context, as just mentioned. Since surface recombination is usually non-radiative and band
bending can lead to the formation of the depletion region where PL is effectively quenched,
both of these phenomena tend to reduce the PL intensity. In the present case, it appears that
silica coating effectively passivates surface dangling bonds and prevents the formation of the

18
described depletion layer. Enlarged portions of the PL spectra, showing transitions from the
different levels in more detail, are shown in Fig. 7.

19
20
21
Fig. 7 Enlarged portions of the spectra showing the Dy3+intraionic transitions in more detail

Dy3+ has a ground electronic configuration [Xe]4f9. The PL emission is dominated by


transitions from the 4F9/2 excited state to the 6H15/2 and 6H13/2 states, giving rise to blue (470–
22
500 nm) and yellow (560–590 nm) PL emissions, respectively. 4F9/2→6H15/2 transitions are of
magnetic dipole (MD) nature, while 4F9/2→ 6H13/2 transitions are of dipole electric origin (ED)
(Gu et al. 2004; Su and Yan 2005; Wang et al. 2009a). All the investigated samples in this
work actually show strong blue and yellow emissions. The dominant peak at 575 nm is due to
the hypersensitive electric dipole transition 4F9/2→6H13/2 (∆J=2) of the Dy3+ ion, which is
strongly influenced by the chemical environment around this rare earth in the host lattice. The
4
F9/2→6H15/2 transition, peaked at 484 nm in the present case, is magnetically allowed and
rarely varies with the crystal field strength and coordination environment around the Dy3+ ion.
The intensity of the yellow emission is much higher than that of the blue one, which
evidences that Dy3+ ions occupy lattice sites without inversion symmetry, in agreement with
Dy3+ replacing Y3+ions at the 4a crystallographic position in D2d symmetry. Actually, the
intensity ratio of the above-mentioned transitions – termed the asymmetry ratio (R) – provides
structural information such as distortion of ligand environment and site symmetry. The R
value was found to be 4.2 for both sol-gel samples, 4.0 for the HT-9 nanopowder, 4.9 for the
HT-2 sample and 4.4 and 7.2 for the HT-2_TEOS_6 and HT-2_TEOS_18 samples,
respectively. The contribution of structural disorder should be considered in order to explain
the different R values, since it might produce different crystal field effects, enhancing the
intensity of the hypersensitive transition due to degeneration of the local symmetry around the
Dy3+ ions. In fact, those samples showing higher asymmetry values, namely samples HT-2
and HT-2_TEOS_18, are those which exhibit narrower peaks in the corresponding Raman
spectra (see Fig. 5).

The differences found in the spectral distribution of the PL emission give rise to different CIE
coordinates, as shown in the chromaticity diagram plotted in Fig. 8. CIE coordinates (x,y)
were calculated to be (0.428, 0.476) and (0.426, 0.472) for SG-2 and SG-9 samples,
respectively, while for HT-2 and HT-9 samples the CIE coordinates were respectively
determined to be (0.436, 0.482) and (0.427, 0.474). In the case of the silica-covered samples,
the calculated CIE coordinates were (0.429, 0.470) and (0.447, 0.486) for HT-2_TEOS_6 and
HT-2_TEOS_18 samples, respectively. The PL emissions of all the investigated samples
correspond then to the yellow region of the visible spectrum. Such spectral characteristics
make Y0.9Dy0.1VO4 a suitable material for applications in white light-emitting diodes (LEDs)
based in the combination of a blue LED chip and a yellow-emitting phosphor.

23
Fig. 8 CIE chromaticity diagram of the Y0.9Dy0.1VO4 and silica-covered samples

Conclusions

Y0.9Dy0.1VO4 nanophosphors were prepared by a sol-gel method and by a hydrothermal


synthesis at acid and basic pH. Silica-covered samples were synthesized from nanoparticles
obtained by the hydrothermal method under acid conditions, as revealed by FTIR and TEM.
According to XRD data, single phases with a zircon-type structure were obtained in all cases.
A higher intensity was observed in XRD patterns of samples synthesized by the sol-gel
process. Size and morphology of the nanoparticles were found to depend on the synthesis
method and preparation conditions, as revealed by TEM images. A clear broadening of the
Raman bands is found in spectra of the samples prepared by the hydrothermal synthesis at pH
= 9, evidencing a certain distortion or non-equivalency of vibrations of VO43- units. The silica
shell of the silica-coated samples was found to be complete and of homogeneous thickness. A
strong PL emission intensity, related to Dy3+ intraionic transitions, was detected in all the
samples investigated. The peak positions of the observed emission lines are not influenced by
the synthesis route followed, which in turn modifies the overall PL intensity as well as the
relative weight of the dominant blue and yellow emissions characteristic of Dy3+ ions. An
enhanced PL emission was found in silica-covered samples, probably due to a reduced
concentration of PL quenching surface defects.
24
Acknowledgements

This work has been supported by MINECO through projects MAT2012-31959, MAT2015-
65274-R, CSD2009-0013 and by Fundación Neurociencias y Envejecimiento through projects
177/2013 and 359/2014.

Conflict of interest

The authors declare that they have no conflict of interest.

25
References

aimay Lin-Vien, Norman B. Colthup WGF and JGG (1991) The Handbook of IR and Raman
Characteristic Frequencies of Organic Molecules
Alcaraz L, Isasi J (2017) Synthesis and study of Y0.9Ln0.1VO4 nanophosphors and
Y0.9Ln0.1VO4@SiO2 luminescent nanocomposites with Ln=Eu, Dy, Er. Ceram Int
43:5311–5318. doi: 10.1016/j.ceramint.2017.01.069
Alcaraz L, Isasi J, Caballero AC, et al (2015a) Nanopowders Y1-yNdyV1-xCrxO4 with y=0 and
1; x = 0, 0.1, 0.2 and 0.5 synthesized by a sol-gel process. Relationship between
morphological characteristics and optical properties. J Lumin 161.: doi:
10.1016/j.jlumin.2014.12.071
Alcaraz L, Isasi J, Díaz-Guerra C, et al (2016a) Preparation of Ca0.5Zr2(PO4)3 and
Ca0.45Eu0.05Zr2(PO4)3 nanopowders: structural characterization and luminescence
emission study. J Phys D Appl Phys 49:115501. doi: 10.1088/0022-3727/49/11/115501
Alcaraz L, Isasi J, Díaz-Guerra C (2015b) Effects of preparation method and pH variation on
the structural characteristics and luminescence properties of Y0.9Er0.1VO4 and
Y0.9Er0.1V0.9Cr0.1O4 nanopowders. J Lumin 165:105–114. doi:
10.1016/j.jlumin.2015.04.038
Alcaraz L, Isasi J, Díaz-Guerra C (2016b) Comparative study of Y0.9Er0.1V1-xPxO4
nanophosphors with x = 0, 0.1, 0.5, 0.9 and 1 prepared by sol-gel and hydrothermal
processes. J Alloys Compd 687:754–764. doi: 10.1016/j.jallcom.2016.06.169
Alcaraz L, Isasi J, Fernández M, Díaz-Guerra C (2014) Effect of synthesis conditions on the
structural characteristics and luminescence properties of Y0.9Eu0.1V1-xCrxO4 (0 ≤ x ≤ 0.5)
nanopowders. Mater Chem Phys 145:18–26. doi: 10.1016/j.matchemphys.2013.12.038
Alcaraz L, Isasi J, Peiteado M, Caballero A (2015c) Síntesis, caracterización estructural y
morfológica de nanofósforos Ca0,45Eu0,05Zr2(PO4)3. Boletín la Soc Española Cerámica y
Vidr 54:236–240. doi: 10.1016/j.bsecv.2015.10.003
Assefa Z, Haire RG, Raison PE (2004) Photoluminescence and Raman studies of Sm 3+ and
Nd3+ ions in zirconia matrices: Example of energy transfer and host-guest interactions.
Spectrochim Acta - Part A Mol Biomol Spectrosc 60:89–95. doi: 10.1016/S1386-
1425(03)00183-5
B. Liu, C.S. Shi ZMQ (2005) Potencial white-light long-lasting phosphors: Dy3+-doped
aluminate. Phys Lett 86:191111
Bao A, Yang H, Tao C, et al (2008) Luminescent properties of nanoparticles YPxV1-xO4:Dy
26
phosphors. J Lumin 128:60–66. doi: 10.1016/j.jlumin.2007.05.011
Cavalli E, Bettinelli M, Belletti A, Speghini A (2002) Optical spectra of yttrium phosphate
and yttrium vanadate single crystals activated with Dy3+. J Alloys Compd 341:107–110.
doi: 10.1016/S0925-8388(02)00079-8
Chen Y, Cheng X, Liu M, et al (2009) Comparison study of the luminescent properties of the
white-light long afterglow phosphors: CaxMgSi2O5+x:Dy3+ (x=1, 2, 3). J Lumin
129:531–535. doi: 10.1016/j.jlumin.2008.12.008
Faoro R, Moglia F, Tonelli M, et al (2009) Energy levels and emission parameters of the Dy3+
ion doped into the YPO4 host lattice. J Phys Condens Matter 21:275501. doi:
10.1088/0953-8984/21/27/275501
Grandhe BK, Bandi VR, Jang K, et al (2012) Multi wall carbon nanotubes assisted synthesis
of YVO4:Eu3+ nanocomposites for display device applications. Compos Part B Eng
43:1192–1195. doi: 10.1016/j.compositesb.2011.08.011
Gu F, Wang SF, Lü MK, et al (2004) Structure Evaluation and Highly Enhanced
Luminescence of Dy3+-Doped ZnO Nanocrystals by Li+ Doping via Combustion
Method. Langmuir 20:3528–3531. doi: 10.1021/la049874f
Hassanzadeh-Tabrizi SA (2012) Synthesis and luminescence properties of YAG:Ce
nanopowder prepared by the Pechini method. Adv Powder Technol 23:324–327. doi:
10.1016/j.apt.2011.04.006
Isasi J, Alcaraz L, Arévalo P, et al (2018) Synthesis and study of (Ca/Ba)0.45Eu0.05Zr2(PO4)3
nanophosphors and (Ca/Ba)0.45Eu0.05Zr2(PO4)3@SiO2 nanostructures with blue-green
emission. J Lumin 204:633–641. doi: 10.1016/j.jlumin.2018.08.082
Jang KH, Sung WK, Kim ES, et al (2009) Time-resolved luminescence spectroscopy of a
YVO4:Eu3+ thin film. J Lumin 129:1853–1856. doi: 10.1016/j.jlumin.2009.01.022
Kai C, Chao G, Bo P, Wei W (2010) The Influence of SiO2 Shell on Fluorescent Properties of
LaF3:Nd3+/SiO2 Core/Shell Nanoparticles. J Nanomater 2010:1–5. doi:
10.1155/2010/238792
Kuang J, Liu Y, Zhang J (2006) White-light-emitting long-lasting phosphorescence in Dy3+-
doped SrSiO3. J Solid State Chem 179:266–269. doi: 10.1016/j.jssc.2005.10.025
Li A-H, Lü Q, Zheng Z-R, et al (2008) Enhanced green upconversion emission of Er3+
through energy transfer by Dy3+ under 800 nm femtosecond-laser excitation. Opt Lett
33:693. doi: 10.1364/OL.33.000693
Li J, Hao Z, Zhang X, et al (2013) Hydrothermal synthesis and upconversion luminescence
27
properties of β-NaGdF4:Yb3+/Tm3+ and β-NaGdF4:Yb3+/Ho3+ submicron crystals with
regular morphologies. J Colloid Interface Sci 392:206–212. doi:
10.1016/j.jcis.2012.09.076
Liang Y, Ouyang J, Wang H, et al (2012) Synthesis and characterization of core-shell
structured SiO2@YVO4:Yb3+,Er33+ microspheres. Appl Surf Sci 258:3689–3694. doi:
10.1016/j.apsusc.2011.12.006
Lin Z, Liu S, Sun X, et al (2014) The effects of citric acid on the synthesis and performance
of silver-tin oxide electrical contact materials. J Alloys Compd 588:30–35. doi:
10.1016/j.jallcom.2013.10.222
Liu B, Kong L, Shi C (2007) White-light long-lasting phosphor Sr2MgSi2O7:Dy3+. J Lumin
122-123:121–124. doi: 10.1016/j.jlumin.2006.01.117
Liu L, Xiao H, An X, et al (2015) Synthesis and photoluminescence properties of core–shell
structured YVO4:Eu3+@SiO2 nanocomposites. Chem Phys Lett 619:169–173. doi:
10.1016/j.cplett.2014.11.065
Liu T, Xu W, Bai X, Song H (2012) Tunable silica shell and its modification on
photoluminescent properties of Y2O3:Eu3+@SiO2 nanocomposites. J Appl Phys
111:064312. doi: 10.1063/1.3694767
Long YW, Yang LX, Yu Y, et al (2006) High-pressure Raman scattering and structural phase
transition in YCrO4. Phys Rev B 74:054110. doi: 10.1103/PhysRevB.74.054110
Luwang MN, Ningthoujam RS, Srivastava SK, Vatsa RK (2011) Preparation of white light
emitting YVO4: Ln3+ and silica-coated YVO4:Ln3+ (Ln3+ = Eu3+, Dy3+, Tm3+)
nanoparticles by CTAB/n-butanol/hexane/water microemulsion route: Energy transfer
and site symmetry studies. J Mater Chem 21:5326. doi: 10.1039/c0jm03470c
Manjón FJ, Rodríguez-Hernández P, Muñoz A, et al (2010) Lattice dynamics of YVO4 at
high pressures. Phys Rev B 81:075202. doi: 10.1103/PhysRevB.81.075202
Miller SA, Caspers HH, Rast HE (1968) Lattice Vibrations of Yttrium Vanadate. Phys Rev
168:964–969. doi: 10.1103/PhysRev.168.964
Moon BK, Kwon IM, Jeong JH, et al (2007) Synthesis and luminescence characteristics of
Eu3+-doped ZrO2 nanoparticles. J Lumin 122-123:855–857. doi:
10.1016/j.jlumin.2006.01.308
Ningthoujam RS, Singh LR, Sudarsan V, Dorendrajit Singh S (2009) Energy transfer process
and optimum emission studies in luminescence of core-shell nanoparticles: YVO4:Eu-
YVO4 and surface state analysis. J Alloys Compd 484:782–789. doi:
28
10.1016/j.jallcom.2009.05.044
Ocaña M, Cantelar E, Cussó F (2011) A facile single-step procedure for the synthesis of
luminescent Ln3+:YVO4 (Ln=Eu or Er+Yb)-silica nanocomposites. Mater Chem Phys
125:224–230. doi: 10.1016/j.matchemphys.2010.09.011
Omkaram I, Buddhudu S (2009) Photoluminescence properties of MgAl2O4:Dy3+ powder
phosphor. Opt Mater (Amst) 32:8–11. doi: 10.1016/j.optmat.2009.05.010
Ray S, Banerjee A, Pramanik P (2009) Shape controlled synthesis, characterization and
photoluminescence properties of YVO4:Dy3+/Eu3+ phosphors. Mater Sci Eng B 156:10–
17. doi: 10.1016/j.mseb.2008.09.049
Sadhu S, Sen T, Patra A (2007) Shape controlled synthesis and luminescence properties of
ZnO:Eu3+ nanostructures. Chem Phys Lett 440:121–124. doi:
10.1016/j.cplett.2007.04.015
Shinde KN, Dhoble SJ, Swart HC, Park K (2012) Phosphate Phosphors for Solid-State
Lighting. Springer Berlin Heidelberg, Berlin, Heidelberg
Shrivastava R, Kaur J, Dubey V (2016) White Light Emission by Dy3+ Doped Phosphor
Matrices: A Short Review. J Fluoresc 26:105–111. doi: 10.1007/s10895-015-1689-8
Shui M, Zheng W, Shu J, et al (2013) Synthesis and electrochemical performance of
Li1+xV3O8 as cathode material prepared by citric acid and tartaric acid assisted sol-gel
processes. Curr Appl Phys 13:517–521. doi: 10.1016/j.cap.2012.09.013
Song WS, Lee KH, Kim YS, Yang H (2012) Tuning of size and luminescence of red
Y(V,P)O4:Eu nanophosphors for their application to transparent panels of plasma
display. Mater Chem Phys 135:51–57. doi: 10.1016/j.matchemphys.2012.04.013
Stober W, Fink A (1968) Controlled Growth of Monodispersed Silica Spheres in the Micron
Size Range. J Colloid Interface Sci 26:62–69
Su XQ, Yan B (2005) The synthesis and luminescence of YPxV1-xO4:Dy3+ microcrystalline
phosphors by in situ co-precipitation composition of hybrid precursors. Mater Chem
Phys 93:552–556. doi: 10.1016/j.matchemphys.2005.04.016
Sun J, Xian J, Xia Z, Du H (2010) Synthesis, structure and luminescence properties of
Y(V,P)O4:Eu3+, Bi3+ phosphors. J Lumin 130:1818–1824. doi:
10.1016/j.jlumin.2010.04.016
Vijayakumar M, Marimuthu K (2015) Structural and luminescence properties of Dy3+ doped
oxyfluoro-borophosphate glasses for lasing materials and white LEDs. J Alloys Compd
629:230–241. doi: 10.1016/j.jallcom.2014.12.214
29
Voron’ko YK, Sobol’ AA, Shukshin VE, et al (2009) Raman spectroscopic study of structural
disordering in YVO4, GdVO4, and CaWO4 crystals. Phys Solid State 51:1886–1893. doi:
10.1134/S1063783409090200
Wang H, Yu M, Lin CK, Lin J (2006) Core-shell structured SiO2@YVO4:Dy3+/Sm3+
phosphor particles: Sol-gel preparation and characterization. J Colloid Interface Sci
300:176–182. doi: 10.1016/j.jcis.2006.03.052
Wang J, Xu Y, Hojamberdiev M, et al (2009a) Optical properties of porous YVO4:Ln (Ln =
Dy3+ and Tm3+) nanoplates obtained by the chemical co-precipitation method. J Alloys
Compd 479:772–776. doi: 10.1016/j.jallcom.2009.01.076
Wang Y, Qin W, Zhang J, et al (2009b) Photoluminescence of colloidal YVO4:Eu/SiO2
core/shell nanocrystals. Opt Commun 282:1148–1153. doi:
10.1016/j.optcom.2008.12.007
Wu C-C, Chen K-B, Lee C-S, et al (2007) Synthesis and VUV Photoluminescence
Characterization of (Y,Gd)(V,P)O4:Eu3+ as a Potential Red-emitting PDP Phosphor.
Chem Mater 19:3278–3285. doi: 10.1021/cm061042a
Xie T, Guo H, Zhang J, et al (2015) Phosphorescence behavior and photoluminescence
mechanism of Dy3+ sensitized β-Zn3(PO4)2: Mn2+ phosphor. J Alloys Compd 642:225–
231. doi: 10.1016/j.jallcom.2015.04.091
Xu H, Wang H, Yan H (2007) Preparation and photocatalytic properties of YVO4
nanopowders. J Hazard Mater 144:82–85. doi: 10.1016/j.jhazmat.2006.09.082
Xue B, Sun J (2014) Upconversion emission properties and tunable morphologies of
Y6WO12:Yb3+/Er3+ phosphor. Infrared Phys Technol 62:45–49. doi:
10.1016/j.infrared.2013.11.001
Y.L. Liu, B.F.Lei CSS (2005) Luminescent properties of a white afterglow phosphor
CdSiO3:Dy3+. Chem Mater 17:2108–2113
Yan B, Su XQ (2005) In situ chemical coprecipitation composition of hybrid precursors to
synthesize YPxV1-xO4:Eu3+ micron crystalline phosphors. Mater Sci Eng B Solid-State
Mater Adv Technol 116:196–201. doi: 10.1016/j.mseb.2004.10.004
You P, Yin G, Chen X, et al (2011) Luminescence properties of Dy3+-doped Li2SrSiO4 for
NUV-excited white LEDs. Opt Mater (Amst) 33:1808–1812. doi:
10.1016/j.optmat.2011.06.018
Yu M, Lin J, Wang Z, et al (2002) Fabrication, Patterning, and Optical Properties of
Nanocrystalline YVO4:A (A = Eu3+ , Dy3+ , Sm3+ , Er3+) Phosphor Films via Sol−Gel
30
Soft Lithography. Chem Mater 14:2224–2231. doi: 10.1021/cm011663y
Yuan H, Wang K, Li S, et al (2013) High-Pressure Stability and Compressibility of Zircon-
Type YV1- x PxO4:Eu3+ Solid-Solution Nanoparticles: An X-ray Diffraction and Raman
Spectroscopy Study. J Phys Chem C 117:18603–18612. doi: 10.1021/jp405405t

31

You might also like